content
stringlengths
1
15.9M
\section{Basic Example} Let $U$ be a local chart of a real 2-dimensional manifold $X$. Take a nondegenerate map $\varphi :U\rightarrow {\Bbb C}^1$ providing $U$ with a complex structure. For each map $f:U\to \bb C$ which is holoorphic with respect to the complex structure above we have the map: $\exp f:U\rightarrow {\Bbb C}^{*}$. The differential of $\exp f$ is a map $A:TU\rightarrow T{\Bbb C}^{*}$, where $T{\Bbb C}^{*}=:g({\Bbb C}^*)=\Bbb C$ \negthinspace is the Lie algebra of the group ${\Bbb C}^{*}$. In the local chart $U\times {\frak g}({\Bbb C}^{*})$ of rank 2 vector bundle the \negthinspace $d(\exp f)$ may be written as the connection form $dy_i+\sum% \limits_{j,k}A_{ij}^ky_jdx_k$ with coefficients \negthinspace $A^1(x_1,x_2)=\left( \begin{array}{cc} u & v \\ -v & u \end{array} \right) \ {\rm and}\ A^2(x_1,x_2)=\left( \begin{array}{cc} v & -u \\ u & v \end{array} \right) $, where $(x_1,x_2)$ are real and imaginary parts of complex structure coordinates induced by $\varphi $, and $(y_1,y_2)$ \negthinspace are coordinates in the fiber ${\frak g}({\Bbb C}^{*})$. This connection is automaticaly flat, since the zero-curvature conditions $\frac{\partial A_i}{\partial x_j}-\frac{% \partial A_j}{\partial x_i}+[A_i,A_j]$\negthinspace $=0$ in this case are just the Cauchy-Riemann conditions \negthinspace $\frac{\partial u}{\partial x_2}% =\frac{\partial v}{\partial x_1},\frac{\partial u}{\partial x_1}=-\frac{% \partial v}{\partial x_2}$. So for a given complex structure we get a family of flat connections corresponding to holomorphic maps. Now vice versa, if we have a flat connection $A$ on a trivial real rank 2 vector bundle over a real 2-dimensional manifold $X$, then, for each local chart $U\times F$ and fixed base point $x'\in X$, we may integrate A along paths $\gamma$ with $\gamma(0)=x'$, wich gives us a map $U\to GL(2,{\Bbb R}),\ x\mapsto G_x:=\int\limits_{\gamma(x',x)}e^A$. If we fix fiber point $y'\in F_{x'}$ then we get a map \negthinspace $\psi :U\rightarrow F\simeq {\Bbb R}^2, x\mapsto G_x(y')$. In particular $\psi(x')=y'$. Now, if we fix a complex structure on $F$ and $A$ is nondegenerate on $U$, then $\psi $ induces a complex structure on $U$. A change of coordinates on $% F$ which preserves complex structure may be represented as an action of the operator \negthinspace $\left( \begin{array}{cc} a & b \\ -b & a \end{array} \right) $. Then a global smooth change of fiber coordinates preserving complex structure on each \negthinspace $F_x$ is the smooth function $\varphi :x\mapsto \left( \begin{array}{cc} a(x) & b(x) \\ -b(x) & a(x) \end{array} \right) $, giving a ''gauge field'' \negthinspace $g(x)=\exp \varphi :X\rightarrow {\Bbb C}^{*}$. Such a change of fiber coordinates induces the transform of connection coefficients \negthinspace $g(A)=A+dg\cdot g^{-1}$. So given a connection \negthinspace $A_0$ we get a map of ''gauge group'' $% {\cal G}$ to the space ${\cal A}$ of connections, such that \negthinspace $% {\rm Id}\mapsto A_0$ and ${\cal A}$ gets fibered into orbits of ${\cal G}$-action. {\it Question:} when a flat connection $A$ may be obtained as the differential of a globally defined map \negthinspace $X\rightarrow F={\Bbb C}^1$, or what is the space ${\cal A}^{\prime }/{\cal G}$ of ${\cal G}$-fibers in the flat connections subspace ${\cal A}^{\prime }\subset {\cal A}$? In particular, when $X$ is orientable and has a fixed complex structure (Riemann surface), then it may be repersented by various connections, which, in general, belong to different ${\cal G}$-fibers. The integration of $A$ with given $v\in x\times F$ in this case is called analytic continuation. \section{General construction} \subsection{Space of Flagged Homotopies} {\it Step 1} Take a $1$-dimensional path $\gamma ^1:[0,1]\rightarrow {\Bbb R}^n$. Its homotopy $\gamma ^2$ with the fixed boundary $\sigma _0=\gamma ^1(0),\sigma _1=\gamma ^1(1)$ sweeps a $2$-disk $D$. The boundary $\partial D=S^1$ gets a decomposition into pairs of cells $((\sigma _0,\sigma _1),(\sigma _0^1,\sigma _1^1))$, where $(\sigma _0,\sigma _1)$ are the boundaries of $\gamma ^1$ (points) and $(\sigma _0^1,\sigma _1^1)$ are the boundaries of the homotopy $% \gamma ^2$ (semi-circles forming $S^1=\partial D$) (see Fig.1). This decomposition will be called {\em flagging} of $D$. For each point $p\in D$ we have two numbers $t_1(p),t_2(p)$ - the values at $p$ of parameters of homotopies $\gamma ^1$ and $\gamma ^2$ correspondingly. Thus we have a coordinatization of $D$. \begin{center} \epsfxsize=15cm Fig.1 \end{center} \noindent {\it Inductive step} Take a $k$-dimensional disk $D^k$ and take its homotopy $\gamma ^{k+1}$ with the fixed boundary $\partial D^k=S^{k-1}$. It sweeps a $k+1$-disk $D$. If $% D^k$ is flagged then $\partial D=S^k$ gets flagging with $k$-cells $(\sigma _0^k,\sigma _1^k)$ being boundaries of $\gamma ^{k+1}$. For each point $p\in D$ we have the value $t_{k+1}(p)$ of parameter of homotopy $\gamma ^{k+1}$ such that $p\in \gamma ^{k+1}(t(p))$. If we fix the homotopy on $\sigma _0^k$, then the corresponding $t^k$-parameter function on $\sigma _0^k$ is extended via homotopy $\gamma ^{k+1}$ to the $t^k$-coordinatization on $D$. So fixing a flag of homotopies $(\gamma ^1,\gamma ^2,\ldots ,\gamma ^k)$ on $(\sigma _0^1,\ldots ,\sigma _0^k)$ correspondingly, where the body $\sigma _0^i$ of each $\gamma ^i=\gamma ^{i+1}(0)$ is the initial point of the subsequent flag element $\gamma ^{i+1}$, gives us values $(t^1(p),\ldots ,t^{k+1}(p))$ of the parameters of these homotopies for each $p\in D$ and thus coordinatization on $D$. For a manifold $X$ the set of homotopies $\gamma ^k$ with fixed flagged boundaries will be denoted by $\Gamma ^k(X)$. In particular, $\Gamma ^0(X)=X$% . For a flagging $\sigma =((\sigma _0,\sigma _1),\ldots ,(\sigma _0^d,\sigma _1^d))$ the set of elements of $\Gamma ^k(X)$ having this flagging is denoted by $\Gamma _\sigma ^k(X)$. Elements of $\Gamma _\sigma ^k(X)$ are subsets of $X$ on one hand and paths in $\Gamma _\sigma ^{k-1}(X)$ on the other. Then for $\gamma ^k\in \Gamma _\sigma ^k(X),\gamma \in \Gamma _\sigma ^{k+1}(X)$ we may write $\gamma ^k\subset \gamma $ or $\gamma ^k\in \gamma $ correspondingly. \subsubsection{Composition Law} Take a pair $\gamma ^{\prime },\gamma ^{\prime \prime }\in \Gamma _\sigma ^k$ such that $\sigma _0^k(\gamma ^{\prime \prime })=\sigma _1^k(\gamma ^{\prime })$. Then we get an element $\gamma ^{\prime \prime }\circ \gamma ^{\prime }\in \Gamma _\sigma ^k$. This composition law makes $\Gamma _\sigma ^k$ into a category called {\em space of flagged homotopies}. \subsubsection{Flagged Homotopy Groups} On the space of flagged homotopies set an equivalence relation: $\gamma ^{\prime }\sim \gamma ^{\prime \prime }$ if $\gamma ^{\prime },\gamma ^{\prime \prime }$ belong to the same connected component of $\Gamma _\sigma ^k$, i.e. there is an element $\gamma ^{k+1}\in \Gamma ^{k+1}$ such that $% (\gamma ^{\prime },\gamma ^{\prime \prime })=(\sigma _0^k,\sigma _1^k)(\gamma )$ (or $\gamma ^{\prime },\gamma ^{\prime \prime }$ are cobordant). Category composition law on $\Gamma _\sigma ^k$ commutes with this factorisation. Take coincident highest dimension cells of $\sigma $, $\sigma _0^{k-1}=\sigma _1^{k-1}$. Then we get a group structure on the corresponding quotient space $\Pi _k(X,\sigma )$. \begin{example} If all boundaries $(\sigma _0^i,\sigma _1^i)$ of $\sigma $ are set (belong) to a single point $p$, then $\Pi _k(X,p)$ becomes Abelian for $k>1$, and $\Pi _k(X,p)=\pi _k(X,p)$ is just the usual $k$-th homotopy group of manifold $X $. \end{example} \subsection{Differential Calculus on $\Gamma _\sigma ^k$} \subsubsection{Tangent to the Space of Flagged Homotopies} For $\sigma \in \Gamma ^k\subset X$ elements of the tangent space to $% \Gamma ^k$ at $\sigma $ are restrictions onto $\sigma $ of vector fields on $% X$ and will be denoted $T_\sigma \Gamma ^k$. \subsubsection{Integration over $\Gamma _\sigma ^k$} Take a $1$-form $\varphi $ on $T_\sigma \Gamma ^k$ with values in algebra $% {\frak g}$. For $\gamma \in \Gamma ^k$ the corresponding homotopy parameter $% t$ is a function on $|\gamma |\subset X$. Take the vector field $v_t:=\frac{% \partial \ln x}{\partial t}:= \lim\limits_{\Delta t\rightarrow 0} \frac{\ln (x_i(t+\Delta t)-x_i(t))}{\Delta t}$. Then we have a map $\int\exp(\varphi):\Gamma ^{k+1}\rightarrow G,\gamma \mapsto {\lim\limits_{\bigtriangleup t\rightarrow 0}} \prod\limits_{i=0}^{N-1}\exp ((\varphi ,v)\Delta t)$, called {\em % multiplicative integral}, where $G=\exp {\frak g}$, is the Lie group, corresponding to Lie algebra of which ${\frak g}$ is the universal envelopping one. The value $\int\limits_\gamma\exp(\varphi)$ of the multiplicative integration functional will also be denoted as $\varphi(\gamma)$. \begin{example} Take a 1-form $\omega \in \Omega ^1(X)$ with values in algebra ${\frak g}$. It is also a $1$-form on $T\Gamma ^0$. For a point $p\subset X$ take a path $% \gamma ([0,1])$ with $\gamma (0)=p$. Take a partition $(t_0,\ldots ,t_N)$ of the interval $[0,1]$. Then we get a value of $\omega $ on $\gamma $ as: $% \int\limits_\gamma\exp(\omega):= {\lim\limits_{\bigtriangleup t\rightarrow 0}} \prod\limits_{i=0}^{N-1}\exp ((\omega ,v(t_i))\cdot (t_{i+1}-t_i))\in G$, where $v(t):=v(x(t))\in T_xX=T_{\gamma (t)}\Gamma ^0$ with the components $v_k(t):= {\lim\limits_{\Delta t\rightarrow 0}} \frac{\ln (x_k(t+\Delta t)-x_k(t))}{\Delta t}$. Notice that the ordering in the ''integral product'' is fixed although the limit does not depend on a choice of parametrisation of $\gamma $. \end{example} \subsubsection{Derivation in $\Gamma _\sigma ^k$} A ${\frak g}$-valued $1$-form $\varphi $ on $T_\gamma \Gamma ^k$ is an operator $T_\gamma \Gamma ^k\rightarrow {\frak g}$. For a pair of tangent vectors $u,v\in T_\gamma \Gamma _\sigma ^k$ (vector fields on $\gamma $) we can take a closed path $\gamma (u,v)$ in $\Gamma _\sigma ^k$ passing through points $\gamma ,\gamma +\exp (u),\gamma +\exp (u)+\exp (v),{\rm and}\gamma +\exp (u)+\exp (v)+\exp (-u)$, where $\gamma +\exp (v)$ is the result of the shift of $\gamma $ along the integral paths of $v$ in $X$. Then take $\varphi ^{\prime }:= {\lim\limits_{\Delta t\rightarrow 0}}\frac{\ln \varphi (\gamma (u\Delta t^1,v\Delta t^2))}{\Delta t^1\Delta t^2}\in {\frak g}$, where $\varphi (\gamma )$ is the multiplicative integral of $\varphi $ over $% \gamma $. This gives us an operator $\wedge ^2T_\gamma \Gamma _\sigma ^k\rightarrow {\frak g}$ called {\em derivative} of $\varphi $. \begin{example} For an algebra ${\frak g}$ and $G=\exp {\frak g}$ take a $G$-valued function $g$ on $\Gamma ^0(X)=X$. Take a path $\gamma ([0,1])\in \Gamma ^1(X)$ with $% \gamma (0)=p_0$ and $\gamma (1)=p_1$ and lets write $g(t):=g(\gamma (t))$. For a partition $(t_0,\ldots ,t_N)$ of the interval $[0,1]$ we have $% g(1)\cdot g(0)^{-1}=\prod\limits_{i=0}^{N-1}g(t_{i+1})\cdot g(t_i)^{-1}=\prod \exp (\frac{\ln (g(t_{i+1})\cdot g(t_i)^{-1})}{\Delta t}% \cdot \Delta t)$. If we define the directional derivative $g_\gamma ^{\prime }:= {\lim\limits_{\Delta t\rightarrow 0}}$ $\frac{\ln (g(t+\Delta t)\cdot g(t)^{-1})}{\Delta t}\in {\frak g}$ then $g(1)\cdot g(0)^{-1}= {\lim\limits_{\Delta t\rightarrow 0}}\prod \exp (g_\gamma ^{\prime }(t_i)\cdot (t_{i+1}-t_i))$. \end{example} \subsection{Variational Calculus on $\Gamma _\sigma ^k$} Take a $1$-form $\varphi \in \Omega ^1(\Gamma ^{k-1})$. It gives us a multiplicative integration functional on $\Gamma ^k$. Take $\gamma \in \Gamma ^k$, and for $u\in T_\gamma \Gamma _\sigma ^k$, take a variation $% \stackrel{\sim }{\gamma }:=\gamma +\delta \gamma =\gamma +\exp (u\Delta t)$ of $\gamma $. Define the {\em first variation} of the functional as $(\delta \varphi ,u):= {\lim\limits_{\Delta t\rightarrow 0}}\frac{\ln (\varphi (\gamma +\exp (u\Delta t))\cdot \varphi (\gamma )^{-1})}{\Delta t}$. It gives us an operator $T_\gamma \Gamma _\sigma ^k\rightarrow {\frak g}$. So, taking variational derivative of a $1$-form on $T\Gamma ^{k-1}$ we get a $1$% -form on $T\Gamma ^k$. \begin{proposition} $(\delta \varphi ,u)= {\lim\limits_{\Delta t\rightarrow 0}}\frac{\ln ((\varphi ^{\prime }\llcorner u\Delta t+\varphi )(\gamma )\cdot \varphi (\gamma )^{-1})}{\Delta t}$. \end{proposition} Take a partition $(t_0,\ldots ,t_N)$ of the domain of the homotopy $\gamma $ and the vector field $v=(\frac{\partial x}{\partial t})$. Then we can write variation of $\varphi $ as an limit of ''integral products'' \begin{eqnarray*} \delta\varphi &=& \prod\limits_{i=0}^{N-1}\varphi(\stackrel{\sim }{\gamma(i)} \stackrel{\sim }\gamma(i+1))\cdot \prod\limits_{i=0}^{N-1}\varphi(\gamma (i+1){\gamma(i)})\\ &=& {\lim\limits_{\Delta t\rightarrow 0}}\prod\limits_{i=0}^{N-1}% \varphi ^{\prime }\llcorner{v(x(i)+\delta x(i))}(t_{i+1}-t_i)\cdot \prod\limits_{i=0}^{N-1}\varphi ^{\prime }\llcorner{v(x(i))}(t_i-t_{i+1})\nonumber \end{eqnarray*} Then we can expand such a product into series in $u_i:=u(x(i))$ as $\delta \varphi =1+f_1\cdot u\Delta t+f_2\cdot u^2\Delta t^2+o(u^2)$, where $f_k\cdot u^k$ denotes the sum of terms of order $k$. In general this is not the expansion of $\exp (f_1\cdot u\Delta t)$ and since the number of commutations in terms of order $\Delta t^2$ has order $\frac 1{\Delta t}$ then these terms make an impact into terms of oreder $\Delta t$. Hence, unlike that in the commutative case, the difference $\varphi (\gamma ^{\prime \prime })\cdot \varphi (\gamma ^{\prime })^{-1}$ is not equal to $ {\lim\limits_{\Delta t\rightarrow 0}} \prod\limits_{j=0}^{N-1}\exp ((\delta \varphi ,u)\cdot \Delta t)$. \subsubsection{Integrable Forms} If the variation of the multiplicative integral functional is identically $0$ then cobordant elements of $\Gamma _\sigma ^k$ have the same value of $% \varphi $, i.e. $\varphi (\gamma )\in G$ is constant on connected components of $\Gamma _\sigma ^k$. Such forms will be called {\em integrable}. Since those components correspond to elements of nonabelian homotopy group $\Pi _k(X,\sigma )$ and the integration functional is a functor to $G$, then we have \begin{Proposition}\label{prp:ifr} Each integrable form gives a representation $\Pi _k(X,\sigma )\to G$. \end{Proposition} The identity $\delta \varphi (u)\equiv 0$ for all $u$ implies that in the series expansion $\delta \varphi =1+(f_1\cdot u)\Delta t+(f_2\cdot u^2)\Delta t^2+\ldots\ $ all the coefficients $f_i=0$. In particular, $f_1\cdot u=% {\lim\limits_{\Delta t\rightarrow 0}}\frac{\ln ((\varphi ^{\prime }\llcorner u\Delta t+\varphi )(\gamma )\cdot \varphi (\gamma )^{-1})}{\Delta t}\equiv 0$ implies $\varphi ^{\prime }(\gamma )\equiv 0$. Then in order to decide on integrability of forms we have to compute their derivatives. The value of the integral of an integrable form in general depends on the flagging $% \sigma $. \begin{example} Take a connection form and an integration path in the form of a torus cut (see Fig.3). For initial point being $A=\sigma _0=\sigma _1$ and clockwise direction the value of the integral is \[ \varphi (\gamma ;A)=aba^{-1}b^{-1} \] For flaging set into point $B$ and the same direction the integral is \[ \varphi (\gamma ;B)=ba^{-1}b^{-1}a \] \begin{center} \epsfxsize=8cm Fig.3 \end{center} \end{example} \subsection{Integration of ${\frak g}$-valued Forms} {\it Step 1} Take a $1$-form $A\in \Omega ^1(X)$ with values in ${\frak g}$ (connection form). Then it is also a $1$-form on $T\Gamma ^0$ so $A$ gives us a measure for multiplicative integration. \noindent {\it Inductive step} Suppose we have defined integration of ${\frak g}$-valued exterior $(k-1)$% -forms on $X$. Take a sequence of forms $(\omega =\omega ^k,\ldots ,\omega ^1)$ where $\omega ^i\in \Omega ^i(X;{\frak g})$. For $\gamma \in \Gamma ^{k-1}$ and $v\in T_\gamma \Gamma _\sigma ^{k-1}$ taking a convolution of $% \omega $ with $v$ we get a $(k-1)$-form $\omega \llcorner v+\omega ^{k-1}\in \Omega ^{k-1}(X)$ on $\gamma \subset X$. Note, that $\omega \llcorner v+\omega ^{k-1}$ in general is not a restriction onto $\gamma $ of any local $(k-1)$-form on $X$. Then we get a $1$-form $\varphi (\omega )$: $v\mapsto {\lim\limits_{\Delta t\rightarrow 0}}\frac{\ln ((\omega \llcorner v\Delta t+\omega ^{k-1})(\gamma )\cdot \omega ^{k-1}(\gamma )^{-1})}{\Delta t% }\in {\frak g}$ on $T_\gamma \Gamma _\sigma ^{k-1}$, also denoted by $\omega $, which is defined by induction. Since computation of $\omega ^{k-1}(\gamma )$ is in its turn dependent of $\omega ^{k-2}$, then $\varphi $ is actually the dependent of the total flag of forms $\varphi =\varphi (\omega ^k,\ldots ,\omega ^1)$. For $\gamma ^k\in \Gamma ^k$ take the vector field $v=(\frac{% \partial x}{\partial t})$ corresponding to the parameter $t^k$ of homotopy. Then we have a map $\Gamma ^k\rightarrow G,\gamma \mapsto {\lim\limits_{\Delta t\rightarrow 0}} \prod\limits_{j=0}^{N-1}\exp (\varphi (\omega ^k,\ldots ,\omega ^1),v\cdot \Delta t)$. Thus each element of $\Omega ^k(X;% {\frak g})$ gives us a measure of integration on $\Gamma _\sigma ^k$. Note, that in general the value of functional $\omega $ depends on $\gamma $. In particular, two homotopies with the same body $|\gamma ^{\prime }|=|\gamma ^{\prime \prime }|\subset X$ but still being different paths in $\Gamma _\sigma ^{k-1}$ may have different values of $\omega $-integral. For each $(k-2)$-dimensional flag $\sigma $ the above defined integration gives a functor form $\Gamma _\sigma ^k$ to $G$. \subsection{Computation of the Exterior Derivative of a ${\frak g}$-valued Form} We have a formal definition of exterior derivative in terms of forms on the space $\Gamma^k$ of paths. In this section we will see \subsubsection{Lattice Approximation of a Homotopy} Take a flag $\gamma =(\gamma ^1,\gamma ^2,\ldots ,\gamma ^k)$ of homotopies and for each $\gamma ^i$ take a partition $(t_0^{(i)},\ldots ,t_{N_i}^{(i)})$ of the interval $[0,1]^{(i)}$. Each $t_j^{(i)}$ defines a hypersurface $% \sigma (t_j^{(i)})$ in $D$, with induced flagging. The set of those hypersurfaces for $i=0,...,k$ induces a cellular decomposition of $D$. For a proper fineness of subdivisions those cells become homeomorphic to cubes. This decomposition will be called {\em lattice approximation} of $D$. For each cube $c_{n_0,\ldots ,n_k}$ take the set of its edges coming out of the ''lowest'' vertex with coordinates $(t_{n_0}^1,\ldots ,t_{n_k}^{(k)})$. An edge with endpoint $(t_{n_0}^1,\ldots ,t_{n_i+1}^{(i)},\ldots ,t_{n_k}^{(k)})$ gives us a vector of $T_{(t_{n_0},\ldots ,t_{n_k})}D$ with coordinates $(0,\ldots ,t_{n_i+1}^{(i)}-t_{n_i}^{(i)},\ldots ,0)$ in the basis $\{\frac \partial {\partial t^{(0)}},\ldots ,\frac \partial {\partial t^{(k)}}\}$. Then the frame of edges at the vertex $(n_0,\ldots ,n_k)$ gives an element of $\wedge ^{k+1}T_{(t_{n_0},\ldots ,t_{n_k}^{(k)})}D$ - the volume element. \begin{proposition} There is a cellular homomorphism $\kappa $ of a $d$-cube onto flagged $d$% -disk. \end{proposition} \begin{example} Let $d=3$. Denote by $v_{ijk},$ $e_{*ij},f_{**i}$ vertices, edges and faces correspondingly. The homomorphism is: \begin{center} $\begin{array}{rl} v_{000}\longmapsto \sigma _0,& v_{111}\longmapsto\sigma_1\\ e_{*00}\cup e_{10*}\cup e_{1*1}\longmapsto \sigma _0^{\prime },& e_{*11} \cup e_{01*}\cup e_{0*0}\longmapsto \sigma _1^{\prime } \\ f_{**0}\cup f_{1**}\cup f_{*1*}\longmapsto \sigma _0^{\prime \prime },& f_{**1} \cup f_{0**}\cup f_{*0*}\longmapsto \sigma _1^{\prime \prime } \end{array}$ \end{center} \end{example} This homomorphism induces a structure of flagged disk on each cubic cell $% c_{n_0,\ldots ,n_k}$ of decomposition of $D^{k+1}$. The homotopy $\gamma ^{(k)}$ may be partitioned into a sequence of infinitesimal homotopies $% \delta \gamma $, where each $\delta \gamma =\delta \gamma _{n_0,\ldots ,n_k}$ is a homotopy, sweeping a cubic cell $c_{n_0,\ldots ,n_k}=|\delta \gamma _{n_0,\ldots ,n_k}|$ and concordant with its flagging structure. This sequence induces an ordering $\Gamma $ on the set of cells. Given a value $% \tau \in \Gamma $ we get a homotopy $\gamma _\tau ^{(k)}=\prod\limits_{\tau ^{\prime }<\tau }\delta \gamma _{\tau ^{\prime }}$ as well as the flag $% (\gamma _\tau ^{(k)},\ldots ,\gamma _\tau ^{(0)})$ of homotopies $\gamma _\tau ^{(i-1)}=\sigma _1(\gamma _\tau ^{(i)})\cap \sigma (t_{n_{i-1}}^{(i-1)})$, where $\sigma _1(\gamma )$ denotes the ending border of homotopy $\gamma $ and $\sigma (t^{(i)})$ denotes the hypersurface with the given value $t^{(i)}$ of $i$-th coordinate on $D$. \begin{center} \epsfxsize=15cm Fig.4 \end{center} For flagged $(d+1)-$disk take its boundary $\partial D=\sigma _0^{(d)}\cup \sigma _1^{(d)}$. The cellular homomorphism $\kappa $ of a cube onto $D$ induces a cellular subdivision of each of $\sigma _0^{(d)}$ and $\sigma _1^{(d)}$ into $d+1$ cubes (see Fig.4). Taking pre-image of the flagging of $% \sigma _i^{(d)}$ induces a flag of chains of faces in $\kappa ^{-1}(\sigma _i^{(d)})$. \begin{proposition} There is a lattice approximation of the homotopy from the subflag $\sigma _0^{(d-1)}$ to $\sigma _1^{(d-1)}$ which may be represented as a ''connected'' path in the set of $d-$cells of $\kappa ^{-1}(\sigma _i^{(d)})$. \end{proposition} Term ''connected'' here means that each next $d-$cell in the sequence has a common $(d-1)-$cell with the previous one. \begin{example} Let $d=2$ and take homomorphism from the previous example. Then the pre-image of flagged 2-sphere $\sigma _0^{\prime \prime }$ consists of $3$ faces $f_{**0},f_{1**},f_{*1*}$. $\partial \sigma _0^{\prime \prime }=\sigma _0^{\prime }\cup \sigma _1^{\prime },\kappa ^{-1}(\sigma _0^{\prime })=f_{*00}+f_{10*}+f_{1*1},\kappa ^{-1}(\sigma _1^{\prime })=f_{0*0}+f_{01*}+f_{*11}$. The deformation from $\kappa ^{-1}(\sigma _0^{\prime })$ to $\kappa ^{-1}(\sigma _1^{\prime })$ is a sequence of chains: $f_{*00}+f_{10*}+f_{1*1},\ f_{*00}+f_{1*0}+f_{11*},\ f_{0*0}+f*_{10}+f_{11*},\ f_{0*0}+f_{01*}+f_{*11}$. We may say that each subsequent chain in this sequence is obtained by sweeping across a $2$-face - the formal ''difference'' between those two chains. For instance $\ f_{*00}+f_{1*0}+f_{11*}=f_{*00}+f_{10*}+f_{1*1}+f_{1**}$. \end{example} \subsubsection{Infinitesimal Loops in $\Gamma _\sigma ^k$} Now take a homotopy of $\sigma _0^{(d-1)}$ to $\sigma _1^{(d-1)}$ across $% \sigma _0^{(d)}$, and then from $\sigma _1^{(d-1)}$ to $\sigma _0^{(d-1)}$ across $\sigma _1^{(d)}$. We get a loop in the space of flagged $(d-1)$% -disks which weeps the surface of $(d+1)$-disk $D$. \begin{example} For $d=3$ the $4$-disk boundary $\partial D=S^3=\sigma _0^{(3)}\cup \sigma _1^{(3)}=D_0^3\cup D_1^3$ consists of two $3$-disks glued along a common sphere $S^2$. In its turn $S^2=$ $\sigma _0^{(2)}\cup \sigma _1^{(2)}=D_0^2\cup D_1^2$ glued along $S^1$. Then we can continuously drag $% D_0^2$ to $D_1^2$ through $D_1^3$ with $S^1$ fixed and then move back to $% D_0^2$ but now through $D_1^3$ thus sweeping the whole $\partial D$. \end{example} For a covering $\kappa $ of $D$ by a cube $c$ the lattice approximation of this loop is an ordered sequence of $d$-faces of the covering cube, denoted by $\overrightarrow{\partial }c$. \noindent {\it Steps 1 and 2} For $\gamma \in \Gamma _\sigma ^k$ denote $\gamma (i):=\gamma (t_i)$. Then for $\gamma ^{\prime },\gamma ^{\prime \prime }\in \Gamma _\sigma ^k$ and a partition $(t_0^k,\ldots ,t_N^k)$ of the domain $[0,1]$ of homotopy we have \begin{eqnarray*} \gamma ^{\prime \prime } &=&(\gamma ^{\prime }(0)\gamma ^{\prime \prime }(0)\gamma ^{\prime \prime }(1)\gamma ^{\prime }(1)\gamma ^{\prime }(0))\cdot (\gamma ^{\prime }(0)\gamma ^{\prime }(1))\cdot (\gamma ^{\prime }(1)\gamma ^{\prime \prime }(1)\gamma ^{\prime \prime }(2)\gamma ^{\prime }(2)\gamma ^{\prime }(1))\cdot (\gamma ^{\prime }(1)\gamma ^{\prime }(2))\cdot \ldots \\ \ &=&\prod\limits_{j=0}^{N-1}(\gamma ^{\prime }(j)\gamma ^{\prime \prime }(j)\gamma ^{\prime \prime }(j+1)\gamma ^{\prime }(j+1)\gamma ^{\prime }(j))\cdot (\gamma ^{\prime }(j)\gamma ^{\prime }(j+1)) \end{eqnarray*} where $(\gamma _1\gamma _2...)$ denotes the path in $\Gamma _{^{(1)}\sigma }^{k-1}$ through a sequence of points $\{\gamma _1,\gamma _2,\ldots \}$. Take a flag of homotopies $(\gamma ^{d+1},\gamma ^d,\ldots ,\gamma ^1)$ and let $\gamma ^{\prime }=\gamma ^{d+1}(i),\gamma ^{\prime \prime }=\gamma ^{d+1}(i+1)$. Then we can write a lattice approximation of the above decomposition as $\gamma (i+1)=\prod\limits_{j=0}^{N-1}\overrightarrow{% \partial }([i,j][i+1,j+1])\cdot ([i,j][i,j+1])$, where $([i,j][i+1,j+1])$ denotes a $2$-cube (square) based on verticies $(i,j)$ and $(i+1,j+1)$, and $% ([i,j][i,j+1]):=(\gamma ^{\prime }(j)\gamma ^{\prime }(j+1))$. For a partition $(t_0^{d-1},\ldots ,t_{N_{d-1}}^{d-1})$ of the domain of $\gamma ^{d-1}$ the $1$-loop $\overrightarrow{\partial }([i,j][i+1,j+1])$ in $\Gamma _{^{(1)}\sigma }^{d-1}$ may be decomposed into a product: $\overrightarrow{% \partial }([i,j][i+1,j+1])=\prod\limits_{k=0}^{N_{d-1}-1}\overrightarrow{% \partial }([i,j,k][i+1,j+1,k+1])\cdot ([i,j,k][i,j+1,k+1])\prod\limits_{k=0}^{N_{d-1}-1}([i,j+1,k+1][i,j,k])$. The sequence $\overrightarrow{\partial }([i,j,k][i+1,j+1,k+1])$ of cube faces is a $2$-loop in $\Gamma _{^{(2)}\sigma }^{d-2}$. \noindent {\it Inductive step} A lattice approximation of $k$-loop $\overrightarrow{\partial }% ([i^{d+1},\ldots ,i^{d-k+1}][i^{d+1}+1,\ldots ,i^{d-k+1}+1])$ in $\Gamma _{^{(k)}\sigma }^{d-k}$ may be decomposed into a product $\begin{array}{l} \prod\limits_{i^{d-k}=0}^{N_{d-k}-1}\overrightarrow{\partial }% ([i^{d+1},\ldots ,i^{d-k}][i^{d+1}+1,\ldots ,i^{d-k}+1])\cdot ([i^{d+1},i^d,\ldots ,i^{d-k}][i^{d+1},i^d+1,\ldots ,i^{d-k}+1])\\ \cdot\prod\limits_{i^{d-k}=0}^{N_{d-k}-1}([i^{d+1},i^d+1,\ldots ,i^{d-k}+1][i^{d+1},i^d,\ldots ,i^{d-k}]) \end{array}$ which is a product of $(k+1)$-paths in $\Gamma _{^{(k+1)}\sigma }^{d-(k+1)}$% . The last term of this sequence of decompositions is $\begin{array}{l} \prod\limits_{i^1=1}^{N_1-1}\overrightarrow{\partial }([i^{d+1},\ldots ,i^1][i^{d+1}+1,\ldots ,i^1+1])\cdot ([i^{d+1},i^d,\ldots ,i^1][i^{d+1},i^d+1,\ldots ,i^1+1])\\ \prod\limits_{i^1=1}^{N_1-1}([i^{d+1},i^d+1,\ldots ,i^1+1][i^{d+1},i^d,\ldots ,i^1]) \end{array}$ \noindent which is a product of $d$-paths in $\Gamma ^0=X$. \subsubsection{Variation of ${\frak g}$-valued Forms} Now having a lattice approximation of the flag of homotopies $(\gamma ^{d+1},\gamma ^d,\ldots ,\gamma ^1)$ we can write $\gamma ^{d+1}(1)\cdot \gamma ^{d+1}(0)^{-1}$ as a path through $d$-faces of $(d+1)$-cubes of the lattice. For a degree $d$ flag of forms $\omega=(\omega^d,\dots,\omega^1)$ we have the value of multiplicative integral functional \begin{eqnarray*} &&\omega (\gamma ^{d+1}(1)\cdot \gamma ^{d+1}(0)^{-1}) =\prod_{i^{d+1}=1}^{N_{d+1}-1}(\prod_{i^d=1}^{N_d-1}\ldots \\ &&\prod_{i^1=1}^{N_1-1}\omega (\overrightarrow{\partial }([i^{d+1},\ldots ,i^1][i^{d+1}+1,\ldots ,i^1+1]))\cdot \omega ([i^{d+1},i^d,\ldots ,i^1][i^{d+1},i^d+1,\ldots ,i^1+1])\\ &&\omega ([i^{d+1},i^d+1,\ldots,i^1+1][i^{d+1},i^d,\ldots ,i^1]) \prod_{i^1=1}^{N_1-1} \omega([i^{d+1},i^d,\ldots ,i^1][i^{d+1},i^d+1,\ldots ,i^1+1])\\ &&\ldots \prod_{i^d=1}^{N_d-1}\omega ([i^{d+1},i^d+1][i^{d+1},i^d])) \end{eqnarray*} where the value of $\omega$ on cells of dimension $k<d$ is the value of the subflag $\omega^{(k)}:=(\omega^k,\dots,\omega^1)$ of corresponding dimension, which is defined by induction. The integral has the following properties: \begin{itemize} \item if the highest degree form $\omega^d=0$ then $\int\exp\omega=1=\exp 0\in\exp g$ \item for a composition of integration paths $\gamma=\gamma'\circ\gamma''$ the value of integral is $\int_\gamma\exp\omega=\int_{\gamma'}\exp\omega\cdot\int_{\gamma''}\exp\omega$ ({\bf multiplicativity}) \item if $\omega^d=d\omega^{(d-1)}$ then $\int_\gamma\exp\omega=\int_{\gamma(1)}\exp\omega^{(d-1)}\cdot(\int_{\gamma(0)}\exp\omega^{(d-1)})^{-1}= \int_{d\gamma}\exp\omega^{(d-1)}$ ({\bf Stokes formula}) \end{itemize} If $\delta _X\omega :=\omega (\overrightarrow{\partial }([i^{d+1},\ldots ,i^1][i^{d+1}+1,\ldots ,i^1+1]))\equiv 0$ then this product contracts to $1$% , then $\omega (\gamma ^{d+1}(1))\equiv \omega (\gamma ^{d+1}(0))$ (i.e. $% \omega $ is integrable) and since for integrable $\omega $ this identity does not depend on the choice of $\gamma ^{d+1}$ then the condition $\delta _X\omega \equiv 0$ is also nesessary and $\delta _X\omega $ is called $X$-% {\em variation} of $\omega $. Each cube $([i^{d+1},\ldots ,i^1][i^{d+1}+1,\ldots ,i^1+1])$ gives us a frame of tangent vectors $% (v_{t_1}\Delta t_1,\ldots ,v_{t_{d+1}}\Delta t_{d+1})$ such that $% x(i^{d+1},\ldots ,i^1)+\exp (v_{t_k}\Delta t_k)=x(i^{d+1},\ldots ,i^k,\ldots ,i^1)$. Then we have an expansion of $\delta \omega $ into series of $% (v_{t_i})$ and the integrability $\delta \omega \equiv 0$ is equivalent to vanishing of coefficients of all terms of this expansion. In particular, the first (homogenious) degree terms give a map $v_{t_1},\ldots ,v_{t_{d+1}}\mapsto {\lim\limits_{\Delta t\rightarrow 0}}\frac{\ln \omega (\overrightarrow{\partial }([i^{d+1},\ldots ,i^1][i^{d+1}+1,\ldots ,i^1+1]))}{\Delta t_1\ldots \Delta t_{d+1}}$ which is a ${\frak g}$-valued $% (d+1)$-form $d\omega \in \Omega ^{d+1}(X;{\frak g})$ called {\em (first) exterior derivative} of $\omega $. Coefficients of monomials of degree $k$ will be called {\em exterior derivatives of order} $k$. \begin{proposition}\label{prp:dfie} A derivative of a form is integrable form. \end{proposition} \begin{example} $k=1$, $\omega =A$ is a connection form of trivial ${\frak g}$-bundle. Then the $1$-loop is a square spanned by line elements $u,v$. The corresponding ''Riemann product'' approximation to the value of Wilson loop $A(\gamma )$ is $e^{A_u(0,0)\cdot u}\cdot e^{A_v(u,0)\cdot v}\cdot e^{-A_u(u,v)\cdot u}\cdot e^{-A_v(0,v)\cdot v}$, where $A_u$ is the component of the connection form in $u$ direction. Then the first derivative \begin{eqnarray*} &&d\omega := {\lim\limits_{u,v\rightarrow 0}}\frac{\ln A(\gamma )}{u\cdot v}% = {\lim\limits_{u,v\rightarrow 0}}\frac{\ln (e^{A_u(0,0)\cdot u}\cdot e^{A_v(u,0)\cdot v}\cdot e^{-A_u(u,v)\cdot u}\cdot e^{-A_v(0,v)\cdot v})} {u\cdot v}\\ &&\ = {\lim\limits_{u,v\rightarrow 0}}\frac{1}{u\cdot v}\ln ((1+A_uu+\frac{(A_uu)^2}{2})(1+(A_v+% \frac{\partial A_v}{\partial u}u)v+\frac{(A_vv)^2}{2})\\ &&\ \ \ \ \ \ \cdot(1-(A_u+\frac{\partial A_u}{% \partial u}u-\frac{\partial A_u}{\partial v}v)u+\frac{(A_uu)^2}{2})(1-(A_v-\frac{% \partial A_v}{\partial u}u)v+\frac{(A_vv)^2}{2}))\\ &&\ = {\lim\limits_{u,v\rightarrow 0}}\frac{\ln (1+(\frac{\partial A_u}{% \partial v}-\frac{\partial A_v}{\partial u}+[A_u,A_v])u\cdot v)}{u\cdot v}=% \frac{\partial A_u}{\partial v}-\frac{\partial A_v}{\partial u}+[A_u,A_v] \end{eqnarray*} which is the curvature form of $A$. The derivatives of higher order are dependents of the first one, so the integrability of $1$-form is equivalent to its being flat. \end{example} \begin{center} \epsfxsize=12cm Fig.5 \end{center} \begin{example} $k=2$, $(\omega ^2,\omega ^1)=(\omega ,A)$ is the $2$-flag of forms. The lattice approximation of an infinitesimal $2$-loop is a cube spanned by line elements $u,v,w$. This loop consists of 6 moves, each sweepping a cube face (see Fig.5). The corresponding ''Riemann product'' approximation to the integral $% \delta \omega $ over the surface of this cube is: \begin{eqnarray*} &&\omega(\overrightarrow{\partial }([0,0,0][u,v,w]))=\\ &&\omega([0,0,0][u,v,0])\cdot A([0][u])\cdot \omega([u,0,0][u,v,w])\cdot A([u][0])\cdot\omega([0,0,0][u,0,w])\cdot A([0][w])\\ &&\cdot \omega([0,0,w][u,v,w])\cdot A([w][0])\cdot \omega([0,0,0][0,v,w])\cdot A([0][v])\cdot\omega([0,v,0][u,v,w])\cdot A([v][0]) \end{eqnarray*} where each term of the product corresponds to moving across one face of the cube in the lattice approximation of the loop $2$-homotopy. Now, in order to find the exterior derivatives of $\omega $ take expansions: $\begin{array}{l} \begin{array}{rl} A([u][0]):=A([u,0,0][0,0,0])=& 1-\sum_i A_iu_i-\sum_{ij}\frac{\partial A_i}{\partial x_j}u_ju_i+\frac{(\sum_iA_iu_i)^2}2+O_3(u)\\ \omega([0,0,w][u,v,w])=& \end{array}\\ \ \ 1+\sum_{ij}\omega_{ij}u_iv_j +\frac{1}{2}\sum_{ijk}\frac{\partial\omega_{ij}}{\partial x_k}u_iv_jw_k +\frac{1}{6}\sum_{ijkl}\frac{\partial^2\omega}{\partial w_k\partial w_l}u_iv_jw_kw_l +\frac{(\sum_{ij}\omega_{ij}u_iv_j)^2}2+O_5(u,v,w) \end{array}$\\ and so on. Then we have expansion into $u,v,w$-series of variation: \begin{eqnarray*} &&\delta \omega =1+\sum_{ijk}u_iv_jw_k(\frac{\partial \omega _{ij}}{\partial x_k}+\frac{% \partial \omega _{ki}}{\partial x_j}+\frac{\partial \omega _{jk}}{\partial x_i}% -([\omega _{ij},A_k]+[\omega _{ki},A_j]+[\omega _{jk},A_i])) \\ &&\ \ +\sum_{ijkl}u_iu_jv_kw_l( \frac \partial {\partial x_i}(\frac{\partial \omega _{jk}}{\partial x_l}+\frac{\partial \omega _{lj}}{\partial x_k}+\frac{\partial \omega_{kl}}{\partial x_j}) -[\omega _{kl},\frac{\partial A_i}{\partial x_j}] +\frac{1}{2}[A_i,\frac{\partial\omega_{kl}}{\partial x_j}]\\ &&\ \ +[\omega_{il},\omega_{jk}]+\frac{\omega_{kl}A_iA_j-2A_i\omega _{kl}A_j+A_iA_j\omega_{kl}}2) \\ &&\ \ +\sum_{ijkl}u_iv_jv_kw_l(\dots)+\sum_{ijkl}u_iv_jw_kw_l(\dots)\\ &&\ \ +O_5(u,v,w) \end{eqnarray*} If we take $\omega =F(A)$ to be the curvature form of $A$, then the vanishing of the coefficient at $uvw$ becomes the Bianchi identity. Note, that vanishing of coefficients of terms of degree 4 does not follow from vanishing of that for $(uvw)$-term, but is satisfied for curvature $F(A)$, so these are higher analogues of Bianchi identity. In commutative case the coefficients of degree 4 reduce to $\frac \partial {\partial x_i}(\frac{\partial \omega _{jk}}{\partial x_l}+\frac{\partial \omega _{lj}}{\partial x_k}+\frac{\partial \omega_{kl}}{\partial x_j})$, so their vanishing follows from vanishing of coefficients of degree 3. This is why the integrability of usual differential forms may be expressed as $d\omega=0$. \end{example} \subsection{Nonabelian Cohomology} \begin{proposition}\label{prp:epr} For integrable local section $\omega\in\Omega^{d+1}(U;{\frak g})$ of the sheaf of $(d+1)$-forms there is a primitive flag $\omega^{(d)}=(\omega^d,\dots,\omega^0)$ of sections $\omega^k\in\Omega^k(U;{\frak g})$ such that $d\omega^{(d)}=\omega^{d+1}$. \end{proposition} Take a local primitive $\omega_U^d\in \Omega ^d(U;{\frak g})$ of global integrable $\omega \in \Omega ^{d+1}(X;{\frak g})$. For a path $s$, with $% s(0)\in U$ take a sequence of open disks $\{U_i\}$ covering $s$. Then we have a sequence of concordant local primitives $\omega _i^d\in \Omega ^d(U_i;% {\frak g})$ of $\omega $, called {\em (analitic) continuation} of $\omega _U^d$, which are concordant on intersections: $\omega _i^d=\omega _{i+1}^d$ on $U_i\cap U_{i+1}$. This gives a functor from $\Gamma ^1(X)$ to the category of local sections of $\Omega ^d(X;{\frak g})$, called {\em complete }$d$-{\em form} (or ''multivalued form''). The space of complete $d$-forms is denoted by $\overline{\Omega ^d(X;{\frak g})}$. If $s$ is a loop then we get another section $\omega _{U+s}^d\in \Omega ^d(U_i;{\frak g})$ on $U$, which is the result of continuation of $\omega _U^d$ along $s$. Since $% d\omega _{U+s}^d=d\omega _U^d$ then $\delta (\omega _{U+s}^d-\omega _U^d)\equiv 0$, and according to Proposition ~\ref{prp:epr} $\omega _{U+s}^d-\omega _U^d=d\omega _U^{d-1}(s)$ for some $\omega _U^{d-1}\in \Omega ^{d-1}(X;% {\frak g})$. Then for each integrable $(d+1)$-form we have a multi-valued (complete) primitive $d$-form. The analytic continuation along a closed loop of the operation of taking local primitive in general sets us onto another branch of the multivalued primitive. Vise versa the derivative of a complete $d$% -form is an integrable univalued $(d+1)$-form, called {\em closed}. The derivatives of single-valued forms (global sections of $\Omega ^d(X;{\frak g})$) will be called {\em exact}. According to Proposition ~\ref{prp:ifr} we have a map $\overline{\Omega ^d(X;{\frak g})}\to Hom(\Pi_{d+1}(X),G)$. {\it Question} If $\omega^i\mapsto\omega^i+\delta\omega^i$ then what is the rule of changing highest degree form $\omega^d$ in order for the integral to transform as $\int_\gamma\exp{\bar\omega}\mapsto \delta(\int_{\gamma^i}\exp{\bar\omega}^i)\cdot \int_\gamma\exp{\bar\omega}\cdot (\delta(\int_{\gamma^i}\exp{\bar\omega}^i))^{-1}$, where $\gamma^i$ is the $0$-th border of the dimension $i$ subflag of the flagged integration cycle $\gamma$. This transform gives us an action of the space $\Omega^i$ of lower degree form on the space $\Omega^d$ of highest order form, called {\em gauge action}. \begin{Definition}\rm The space of orbits of gauge action is called the {\em space of cohomologies} with coefficients in ${\frak g}$, denoted $H^{d+1|k}(X;{\frak g})$. \end{Definition} \begin{example} Take $X=S^1$,${\frak g}={\Bbb R}$, $d=0$. Then a complete $0$-form is a multi-valued function $f$ with $d(f_{U+s}-f_U)=0$, and hence the ''discrepancy'' $\Delta :=f_{U+s}-f_U$ is constant. Adding a global section to a comlete $0$-form preserves $\Delta $. And vise versa the difference of two complete $0$-forms with the same discrepancy is a single-valued function. So the space of gauge orbits of complete $0$-forms is parametrized by values of $\Delta $, and then $H^{1|0}(S^1,{\Bbb R})={\Bbb R}^1$. \end{example} \begin{example} In the framework of Section 1 take $X={\bb R}^2-\{0\}$. For a flat $g({\Bbb C}^*)$-valued connection form (or a locally holomorphic function on $X$) we get the value of monodromy integral $\int_\gamma e^{A}\in {\Bbb C}^*$. Because of commutativity of algebra $g({\Bbb C}^*)$ the gauge action reduces to $A\mapsto A+df$, where $f$ is a globally defined smooth $g({\Bbb C}^*)$-valued function. This action preserves the monodromy, but in general moves $A$ to a different complex structure class. Restricting this action to some fixed comlex structure we get the space of gauge orbits ${\cal A'}/{\cal G}=H^{1|0}({\bb R^2}-\{0\},g({\Bbb C}^*)) =\{z^\alpha|\alpha\in{\bb C}\}/\{z^n|n\in\bb Z\}={\bb C}/{\bb Z}$. \end{example} \section{Applications} \subsection{De Rham theory} Here $G={\Bbb R}_{+}^{*}$, ${\frak g=}{\Bbb R}^1$. Since $G$ is commutative we can write all formulas in additive notations taking the logarithm of corresponding multiplicative ones. Then in $\delta \omega $ the first derivative is the usual exterior derivative $d\omega $ of a form and vanishing of $d\omega $ implies vanishing of all higher derivatives. Thus integrability is equivalent to $d\omega =0$. Since the derivative of a form is integrable this implies the identity $d^2\omega \equiv 0$. The space of closed $(d+1)$-forms is in 1-1 coorrespondence with complete $d$-forms (Poicare Lemma). Then the space $H^{d+1|d}(X):=\overline{\Omega ^d(X;{\Bbb R}% )}/\Omega ^d(X;{\Bbb R})$ of orbits of the action of $\Omega ^d(X;{\Bbb R})$ on complete $d$-forms corresponds to the De Rham cohomology space $H_{DR}^d(X;% {\Bbb R})$. Because of commutativity the gauge action is a morphism of affine spaces. Then the identity $d^2\omega \equiv 0$ implies that the action of $% \Omega ^k(X;{\Bbb R})$ on $\overline{\Omega ^d(X;{\Bbb R})}\subset \Omega ^{d+1}(X;{\Bbb R})$ is trivial for $k<d$, so $H^{d+1|k}(X)=\overline{% \Omega ^d(X;{\Bbb R})}$. \subsection{Gauge theory} Here $d=0$. The space of global $0$-forms is the space of $G$-valued functions. It has a group structure called gauge group. Its action on the space $\Omega ^1(X;{\frak g})$ of connections is given by the formula $% g:A\mapsto gAg^{-1}+g^{-1}dg$. The space $\overline{\Omega ^0(X;{\frak g})}$ is in 1-1 correspondence with the space of flat connections. Since $% d(g_{U+s}-g_U)=0$ then the discrepancy $\Delta (s):=g_{U+s}-g_U$ is constant on $X$ similar to commutative case. A representative of a cohomology class of $g\in \overline{\Omega ^0(X;{\frak g})}$ (or of a flat connection) is given by the set of values $\Delta (s)$ on loops $s\in \Gamma _x^1$, where $x=s(0)$, or, since $\Delta (s)$ depends only on homotopy class of $s$, by the representation $\pi _1(X)\rightarrow G$. Since the gauge group acts on $\Delta(s)$ by conjugation then $H^{1|0}(X)={\rm Hom}(\pi_1(X),G)/G$ is the space of conjugacy classes of representations $\{\pi_1(X)\to G\}$. The consequence $ddA\equiv 0$ of Proposition ~\ref{prp:dfie} is known as Bianchi identity. Note that here the action of the gauge group does not commute with the affine space structure. In particular the action of $\Omega ^0(X;{\frak g})$ on the space $d\Omega ^1(X;{\frak g})$ of curvature forms is not trivial and given by $g:F\mapsto gFg^{-1}$. Then the space of orbits $H^{2|0}(X;{\frak g}):=\overline{\Omega ^1(X;{\frak g})}% /\Omega ^0(X;{\frak g})\neq \overline{\Omega ^1(X;{\frak g})}$ unlike that in the commutative case. \flushleft \vspace{0.2 cm} Author address:\\ College of Mathematics,\\ Independent University of Moscow \end{document}
\section{\bf Introduction} Integrable models, since its inception, occupy a distinguised position because of their striking predictability and the existence of a special class of extended solutions, called solitons. Solitons are characterised by non-dispersive localised wave pulses which produce as a result of the competetion between linear dispersion and nonlinearity. Moreover, solitons retain thier shapes even after collisions among themselves and consequently exhibit particle like behaviour. Integrable models as well as solitons have applications in such diverse areas of physics as high energy physics, condensed matter physics, plasma physics, nonlinear optics and nuclear physics. At present, a handful of nonlinear dynamical equations exist exhibiting soliton solutions. Historically, KdV equation \cite {1} is the most important one. It describes the dynamics of waves moving in shallow water. KdV and KP hierarchies also play an important role in describing non-critical strings \cite {2} through the construction of $\tau$ functions. In KdV or KP hierarchy approach, evaluation of partition function and all correlation functions may be made directly through a set of integrable differential equations, bypassing the complicated technicalities involved in other approaches. Sine-Gordon equation \cite {3} is another welknown example of integrable equation, which first arose as a master equation for pseudo-spherical surfaces. Besides, it describes the propagation of magnetic flux in a Joseption junction transmission line, the propagation of dislocations in a crystal lattice and many other physical problems. Nonlinear Schr{\" o}dinger equation and its higher order generalizations [4,5] have also many applications in plasma physics and in nonlinear optics. In particular recently, a lot of excitement is veered around the soliton solutions of a higher order generalization of nonlinear Schr{\" o}dinger equation, because of its potential application in high speed fibre communication systems. The key observation is that the dynamics of the higher order nonlinear Schr{\" o}dinger equation (HNLS) \cite {6} not only takes care of dispersion loss, but also takes care of the propagation loss as the optical pulse propagates along the fibre. This is due to the fact that stimulated Raman effect \cite {6a}, which compensates the propagation loss, already exists withing the spectrum of HNLS equation. As a consequence, a very short pulse can be propagetd over a long distance without distortion. Thus, it is optical solitons, which are responsible behind the successful propagation of optical pulses through the fibre over a long distance. Therefore, the importance of the study of integrable models and to find their soliton solutions is unquestionable. There are several methods exist to obtain soliton solutions of a nonlinear evolution equation, like Hirota bilinear method \cite {7}, Painleve analysis \cite {8}, B{\" a}cklund transformation \cite {9} and inverse scattering transform (IST) \cite {10}. IST method, however, is the most elegant tool, which eventually proves the complete integrability of the evolution equation. IST method, on the other hand, is the most complicated one and is intimately connected with the existence of an auxiliary linear problem, known as Lax equations. Most of the known integrable evolution equations are, in fact, associated with $2\times 2$ and $3\times 3$ dimensional Lax operators. But there are some evolution equations whose associated linear equations, in general, cannot be cast in terms of $2\times 2$ or $3\times 3$ Lax operators. One such example is higher order nonlinear Schr{\" o}dinger equation (HNLS) \cite {6}: \begin{equation} i\partial_t q + {\beta }_1 \partial_{xx} q + {\beta }_2 \vert q\vert^2 q i\epsilon [\beta_3 \partial_{xxx}q + \beta_4 \partial_x(\vert q\vert^2)q + \beta_5 \vert q\vert^2\partial_x q] = 0 \label{1.1} \end{equation} where, $q$ is a complex field and $\beta_i\quad i=1,2\cdots,5$ are the constant coefficients. Notice that in absense of last three terms, (\ref {1.1}) reduces to nonlinear Schr{\" o}dinger equation. It is observed recently, for arbitrary values of the coefficients, (\ref {1.1}) is associated with $n\times n$ Lax operators \cite {11}. Interestingly, the parameters, $\beta$ are related to the dimension of the Lax operators. Unfortunately, IST method is not well understood for arbitrary dimensional Lax oparators. Nonetheless, IST methods developed in the context of $2\times 2$ \cite {10} and $3\times 3$ \cite {12,12a} matrix Lax operators cannot be generalized straight forwardly to handle more general cases. Thus, in order to obtain soliton solutions of (\ref {1.1}) by IST method, a generalized IST method, associated with $n$ dimensional Lax operator, is to be formulated. In this paper, we will, therefore, develop a generalized IST method, which ultimately leads to solving a set of coupled Gelfand Levitan Marchenko equations \cite {12b} for obtaining soliton solutions. We will use the generalized IST method to obtain soliton solutions for an evolution equation, whose dynamical field is an $(n-1)$ component vector, $\vec q$. $\vec q$, in fact, is a vector generalization of the field $q$ given in (\ref {1.1}). The evolution equation may, therefore, be called vector higher order nonlinear Schr{\" o}dinger equation (VHNLS) and will be described in the next section. We will also show that under suitable reduction VHNLS equation reduces to (\ref {1.1}). The rigidity in the structures of the solitons reveals that the system has a huge underlying symmetry. This symmetry is, in general, manifested by the existence of an infinite number of conserved quantities. Existence of infinite number of conserved quantities is a key criterion for a field theoretical model to be called integrable in Liouville sense \cite {13}. Construction of Lax pair of a dynamical system although itself is a good hint for integrability, many integrable models may not satisfy Liouville integrability criterion. In this paper we will obtain Riccati equations associated with an $n$ dimensional Lax oparator and indentify the generating function of the conserved charges. Consequently, we will find conserved charges explicitly in terms of the field variables and their derivatives. Another interesting aspect, we will address, is to establish a connection between nonlinear field theories and spin systems. In this context, it is welknown that nonlinear Schr{\" o}dinger equation is gauge related to Landau Lifshitz equation \cite {14}. Recently Sasa Satsuma equation \cite {14a}, a particaluar version of HNLS equation is shown to be gauge related to a generalized Landau Lifshitz equation, where the spin field is associated with $SU(3)$ group \cite {15,16}. It is thus expected that a spin system may also exist corresponding to a vector generalisation of HNLS equation. We will establish a connection between the VHNLS equation, proposed by us and a generalized Landau Lifshitz equation and obtain a Lax pair for the spin system. The organization of the paper is as follows. In section 2, the VHNLS equation is introduced and a Lax pair for the nonlinear equation is constructed. We also show a reduction procedure under which VHNLS equation reduces to a two parameter family of HNLS equation. In section 3, we show the existence of infinite number of conserved quantities, by solving a set of coupled Riccati equations. We develop the inverse scattering method for an $n$ dimensional Lax operators in section 4 and obtain $n$ Gelfand Levitan Marchenko equations. Section 5 deals with the solutions ofGelfand Levitan Marchenko equations and cosequently find $N$ soliton solutions. We show in section 6 the gauge equivalence of VHNLS equation and the generalised Ladau Lifsitz equation and obtain a Lax pair for the generalized Landau Lifshitz equation in terms of the spin field. Section 7 is the concluding one. \section{\bf Evolution Equation and Lax Pair} We propose a vector nonlinear evolution equation of the form \begin{equation} \vec{q}_t + \epsilon \vec{q}_{xxx} + 3\epsilon (\vec{q^*}\cdot\vec{q}_x) \vec{q} + 3\epsilon \vert \vec{q}\vert^2 \vec{q}_x = 0 \label{2.1} \end{equation} where, the dynamical field variable, $\vec{q} = (q_1,q_2,\cdots ,q_{n-1})$, is an $(n-1)$-tuple vector. The equation (\ref {2.1}) is an example of vector higher order nonlinear Schr{\" o}dinger equation (VHNLS). The vector field $\vec q$ may be interpreted as an $(n-1)$ interacting optical modes describing the dynamics of a charge field with $(n-1)$ colours. In fact, we will show under suitable reduction, (\ref {2.1}) yields to the HNLS equation \cite {11}. In order to obtain Lax pair for (\ref {2.1}), we first introduce the $n\times n$ matrix linear eigenvalue problem as \begin{subequations} \begin{eqnarray} \partial_x \Psi &=& {\bf U}(x,t,\lambda)\Psi \\ \label{2.2a} \partial_t \Psi &=& {\bf V}(x,t,\lambda)\Psi \label{2.2b} \end{eqnarray} \label{2.2} \end{subequations} where, ${\bf \Psi }(x,t)$ is an $n$-tuple vector auxilliary field, $\lambda $ is the spectral parameter and the Lax operators ${\bf U}(x,t)$ and ${\bf V}(x,t)$ are $n\times n$ matrices. Let us now assume an explicit form of the Lax pair, ${\bf U}(x,t)$ and ${\bf V}(x,t)$, associated with the VHNLS equation, as \begin{subequations} \begin{eqnarray} {\bf U} &=& -i \lambda {\bf \Sigma} + {\bf A} \\ {\bf V} &=& - \epsilon {\bf A}_{xx} + \epsilon ({\bf A}_x{\bf A} -{\bf A}{\bf A}_x) + 2\epsilon {\bf A}^3 \\ \nonumber & & -2i\epsilon \lambda {\bf \Sigma} ({\bf A}^2 -{\bf A}_x) + 4\epsilon {\lambda}^2{\bf A} -4i\epsilon {\lambda}^3{\bf \Sigma}, \end{eqnarray} \label{2.3} \end{subequations} In the equation (\ref {2.3}) ${\bf \Sigma }$ is a c-no. diagonal matrix and the matrix ${\bf A}(x,t)$ consists of dynamical fields, $\vec{q}(x,t)$ and $\vec{q^*}(x,t)$ only. It is interesting to note that the evolution equation for the matrix ${\bf A}(x,t)$ immediately follows from the compatibility condition, namely \[ {\bf U}_t(x,t) - {\bf V}_x(x,t) + [{\bf U}(x,t) , {\bf V}(x,t)] = 0 \] provided ${\bf A}$ and ${\bf \Sigma}$ satisfy the conditions that \begin{equation} {\bf \Sigma}^2 = {\bf 1}, \quad\quad {\bf \Sigma}{\bf A} + {\bf A}{\bf \Sigma} = 0 \label{2.4} \end{equation} and as a consequence, the nonlinear evolution equation for ${\bf A}(x,t)$ becomes \begin{equation} {\bf A}_t + \epsilon {\bf A}_{xxx} - 3\epsilon ({\bf A}^2 {\bf A}_x + {\bf A}_x {\bf A}^2) = 0 \label{2.5} \end{equation} It is now clear that various representations of the matrix ${\bf A }$ in terms of the fields $\vec{q}$ and $\vec{q^*}$ yield to different nonlinear evolution equations. To associate (\ref {2.5}) with the VHNLS equation (\ref {2.1}), let us consider the explicit expressions of ${\bf \Sigma }$ and ${\bf A}(x,t)$ of the form satisfying the properties (\ref {2.4}) as \begin{subequations} \begin{eqnarray} {\bf \Sigma} &=& \sum_{i=1}^{n-1} e_{ii} - e_{nn} \\ \label{2.6a} {\bf A}(x,t) &=& \sum_{i=1}^{n-1} q_i(x,t) e_{in} - \sum_{i=1}^{n-1} q_i^*(x,t) e_{ni} \label{2.6b} \end{eqnarray} \label{2.6} \end{subequations} where, $e_{ij}$ is an $n\times n$ matrix whose only $(ij)$th element is unity, the rest elements being zero and $q_i(x,t)$ is the $i$th. component of the dynamical field $\vec q$. Substituting (\ref {2.6}) in (\ref {2.5}), the evolution equation becomes \begin{equation} q_{it} + \epsilon q_{ixxx} + 3\epsilon (\sum_{j=0}^{n-1} q_j^*.q_{jx}) q_i + 3\epsilon (\sum_{j=0}^{n-1} q_j^*q_j) q_{ix} = 0 \label{2.7} \end{equation} which is nothing but VHNLS equation (\ref {2.1}), written in the component form. Now in order to obtain HNLS equation (\ref {1.1}) let us consider the following reduction. Notice that all the dynamical fields $q_i$ in (\ref {2.7}) are independent. However, instead of $(n-1)$ independent dynamical fields, if we restrict ourselves to only one dynamical field $q$ and its complex conjugate $q^*$ in (\ref {2.7}), then all the $q_i$s are not independent. $q_i$s, in this case, may be chosen as either $q$ or $q^*$. If we, for example, choose $m$ number of $q_i$ as $q$ and the rest $(n-m-1)$ numbers as $q^*$, (\ref {2.7}) reduces to HNLS equation \begin{equation} q_t + \epsilon q_{xxx} + 6m\epsilon\vert q\vert^2 q_x + 3(n-m-1)\epsilon (\vert q\vert^2)_x q = 0. \label{2.8} \end{equation} The equation (\ref {2.8}), to be precise, is a gauge equivalent version of HNLS equation. Soliton solutions of (\ref {2.8}) and those of HNLS equation are, in fact, related through a U(1) gauge tranformation \cite {11}. Two well known euqations immediately follow from (\ref {2.8}). If we choose $m=n-1$, (\ref {2.8}) gives rise to Hirota equation \cite {17} \begin{equation} q_t + \epsilon q_{xxx} + 6\epsilon\vert q\vert^2 q_x = 0 \label{2.9} \end{equation} after rescalling of $q$ and $q^*$ as $q=(n-1)^{-\frac{1}{2}}q$ and $q^*=(n-1)^{-\frac{1}{2}}q^*$ respectively. On the other hand, if we choose $m=(n-1)/2$, which is only possible for odd dimensional Lax pair, (\ref {2.8}) reduces to Sasa Satsuma euqation \cite {14a,16,18}, \begin{equation} q_t + \epsilon q_{xxx} + 6\epsilon\vert q\vert^2 q_x + 3\epsilon (\vert q\vert^2)_x q = 0, \label{2.10} \end{equation} once again with appropriate scalling of the fields $q$ and $q^*$ respectively as $q=(\frac{n-1}{2})^{-\frac{1}{2}}q$ and $q^*=(\frac{n-1}{2})^{-\frac{1}{2}}q^*$. \vspace{.5 cm} \section {\bf Riccati Equation and Conserved Charges} In order to obtain Riccati equation, we first write the Lax equation (\ref {2.2}a,\ref {2.3}a,\ref {2.6}) in the component form. For the first $(n-1)$ components of ${\bf \Psi}$, (\ref {2.2}a) can be written in the form, \begin{eqnarray*} \Psi_{1x} &=& -i\lambda \Psi_1 + q_1 \Psi_n\\ \Psi_{2x} &=& -i\lambda \Psi_2+ q_2 \Psi_n\\ & &\vdots\\ \Psi_{n-1x} &=& -i\lambda \Psi_{n-1} + q_{n-1} \Psi_n\\ \end{eqnarray*} {\it i.e.} \begin{subequations} \begin{equation} \Psi_{ix} = -i\lambda \Psi_i + q_i \Psi_n \label{3.1} \end{equation} where, $i = 1,2,\cdots , n-1 $ and $\Psi_i$ denotes the $i$th. component of ${\bf \Psi}$. But for the $n$th. component of ${\bf \Psi}$, the Lax equation (\ref {2.2}a) has a different form as \begin{eqnarray} \Psi_{nx} &=& i\lambda \Psi_n - q_1^* \Psi_1 - q_2^* \Psi_2 \cdots - q_{n-1}^* \Psi_{n-1}\nonumber \\ &=& i\lambda\Psi_n - \sum_{j=1}^{n-1}q_j^* \Psi_j \label{3.2} \end{eqnarray} \end{subequations} Following now a similar procedure as in \cite {16} we write, \begin{equation} \Gamma_i = \frac{\Psi_i}{\Psi_n}, \label{3.3} \end{equation} $i = 1,2,\cdots, n-1$, which are related to the conserved charges $\alpha_{nn}(\lambda)$ in the following way, \begin{eqnarray} \ln \alpha_{nn}(\lambda) &=& \ln \Psi_n -i\lambda x\vert_{x\rightarrow \infty}\nonumber \\ &=& -\int_{-\infty}^{\infty}dx (\sum_{i=1}^{n-1}q_i^*\Gamma_i) \label{3.4} \end{eqnarray} We will see in section 5 that $\alpha_{nn}(\lambda)$ is, indeed, $(nn)$th element of the scattering data matrix and more so it does not evolve with time. By using (\ref{3.1},b) and (\ref{3.3}), we may obtain a first order differential equation for each $\Gamma_i$, \begin{equation} \Gamma_{ix} + 2i\lambda\Gamma_i - \sum_{j=1}^{n-1}q_j^*\Gamma_j\Gamma_i - q_i = 0\\ \label{3.5} \end{equation} The set of $(n-1)$ coupled nonlinear differential equations for $\Gamma_i$ in (\ref {3.5}) are called Riccati equations. It is obvious from (\ref {3.4}) that the solutions of Riccati equations eventually determine the conserved quantities. Now in order to solve (\ref{3.5}), we assume a series solution of $\Gamma_i$ as \begin{equation} \Gamma_i(x,\lambda) = \sum_{n=0}^{\infty}C^i_n(x)\lambda^{-n} \label{3.6} \end{equation} Substituting (\ref{3.6}) into (\ref{3.5}), the following recursion relations may be obtained. \begin{subequations} \begin{eqnarray} C_0^i = 0 ;\quad \quad \quad C^i_1 = \frac{q_i}{2i} \end{eqnarray} and \begin{equation} 2iC_{k+2}^i + (C^i_{k+1})_x - \sum_{m=0}^{k-1}C^i_{k-1+m}\sum_{j=1}^{n-1}q_j^*C^j_m =0\\ \label{3.7} \end{equation} \end{subequations} with $ k= 0,1,2,.......... $. The infinite number of Hamiltonians (conserved quantities) may explicitly be determined in terms of the dynamical field variables $q_i$ and their derivatives by expanding $\alpha_{nn}(\lambda)$ in the form, \begin{equation} ln\alpha_{nn}(\lambda) = (n-1)\sum_{l=0}^{\infty}\frac{(-1)^l}{(2i)^{2l +1}} H_l\lambda^{-l} \label{3.8} \end{equation} and thus comparing (\ref {3.8}) with (\ref {3.4}) and (\ref {3.6}), ${H_l}$ becomes \begin{equation} H_l = \frac{(2i)^{2l+1}}{(-1)^l(n-1)}\int dx[\sum_{j=1}^{n-1}q^*_jC^j_l)] \label{3.9} \end{equation} The explicit expresseions of the first few low order Hamiltonians are given below. \begin{subequations} \begin{eqnarray} H_1 &=& \frac{1}{n-1}\int dx [\sum^{n-1}_{j=1}q_j^*q_j] \\ \label{3.10a} H_2 &=& \frac{1}{2(n-1)}\int dx \sum_{j=1}^{n-1}[q_j^*q_{jx} - q^*_{jx}q_j]\\ \label{3.10b} H_3 &=& \frac{1}{n-1}\int dx [(\sum_{j=1}^{n-1}q_j^*q_j)^2 - \sum_{j=1}^{n-1} q_{jx}^*q_{jx}]\\ \label{3.10c} H_4 &=& \frac{1}{2(n-1)}\int~dx\sum_{j=1}^{n-1}[q_j^*q_{jxxx}-q^*_{jxxx}+ 3(\sum_{k=1}^{n-1}\vert q_k\vert)^2(q_j^*q_{jx}-q^*_{jx}q_j)]\nonumber\\ &~&\\ \label{3.10d} H_5 &=& \frac{1}{n-1}\int dx[\sum_{j=1}^{n-1}q_{jxx}^*q_{jxx} + 2(\sum_{j=1}^{n-1}q^*_jq_j)^3 - (\sum_{j=1}^{n-1}(\vert q_j\vert^2)_x)^2 \nonumber \\ &-& 4\sum_{k=1}^{n-1}\vert q_j\vert^2\sum_{j=1}^{n-1}q_{jx}^*q_{jx} - 2\sum_{j=1}^{n-1}q_j^*q_{jx}\sum_{k=1}^{n-1}q^*_{kx}q_k] \label{3.10e} \end{eqnarray} \label{3.10} \end{subequations} We verify directly by using equations of motions that $H_1$, $H_2$, $H_3$, $H_4$ and $H_5$ are, indeed, constants of motions. Let us now specialize to HNLS equation. If we assume that $m$ number of $q_i$s are chosen as $q$ and the rest $n-m-1$ as $q^*$, the Hamiltonians in (\ref {3.10}) reduce to \begin{subequations} \begin{eqnarray} H_1 &=& \int dx \vert q \vert^2\\ \label{3.11a} H_2 &=& (2m -n+1)\int dx (q_xq^* - qq^*_x)\\ \label{3.11b} H_3 &=& \int dx ((n-1)\vert q \vert^4 - q_x q^*_x)\\ \label{3.11c} H_4&=&(2m-n+1)\int~dx [3(n-1)\vert q\vert^2 (q_xq^*-qq^*_x) +\frac{1}{2}(q_{xxx}q^*-qq^*_{xxx})]\nonumber \\ &~&\\ \label{3.11d} H_5&=&\int~dx [q_{xx}q^*_{xx}+2(n-1)^2 \vert q \vert^6 +[(2m-n-1)^2 -4(n-1)]\vert q \vert^2 q_x q^*_x\nonumber \\ &-&[(n-1) + \frac{2m(n-l-1)}{(n-1)}]((\vert q \vert^2)_x)^2]\\ \nonumber &~& \label{3.11e} \end{eqnarray} \label{3.11} \end{subequations} Notice that $H_1$ in (\ref {3.11}a) has a universal form, which implies that all solitons have the same energy irrespective of their shapes. However, the momentum, $H_2$ crucially depends on the numbers of $q$, chosen in a given representation of the matrix ${\bf A}$ (\ref {2.6}). For example, $H_2$ becomes zero for Sasa Satsuma case, {\it i.e.} for $m=(n-1)/2$. In fact, all conserved quantities having even indices become trivial for Sasa Satsuma case. $H_3$ in (\ref {3.11}c) also has a somewhat universal form, depending only on the dimensions of the Lax pairs. But the higher order conserved quantities starting from $H_5$ do not possess such universal forms. Their explicit forms depend both on the dimensionality of the matrix Lax pair and also on the representations of the matrix ${\bf A}$. \section{\bf Generalized Gelfand Levitan Marchenko Equations} We now generalize IST method suitable for studying $n$ dsmensional Lax operators. The first step in this direction is to obtain a set of generalized Gelfand Levitan Marchenko equations. This generalisation is a nontrivial one and crucially depends on the properties of scattering data matrix. However, we will see that for three dimenisonal Lax oparators generalized Gelafand Levitan Marchenko equations will reduce to Sasa Satsuma case \cite {14a} as is expected. We broadly follow the treatment of Manakov, developed in the context of $3\times 3$ Lax operator \cite {12a} and for that assume Jost functions, for real $\lambda$, satisfy the boundary conditions, \begin{subequations} \begin{equation} \Phi^{(i)} = e_i e^{-i\lambda x} \label{IV1a} \end{equation} with $ i=1,2.......n-1$, but the $n$th. one satisfies a different boundary condition like \begin{equation} \Phi^{(n)} = e_n e^{i\lambda x} \label{IV1b} \end{equation} \label{IV1} \end{subequations} as $x\rightarrow -\infty$. Similarly, as $x\rightarrow \infty$, other set of Jost functions satify the boundary conditions \begin{subequations} \begin{equation} \Psi^{(i)} = e_i e^{-i\lambda x} \label{IV2a} \end{equation} for $ i= 1,.......n-1$ and for the $n$th. one, \begin{equation} \Psi^{(n)} = e_n e^{i\lambda x}. \label{IV2b} \end{equation} \label{IV2} \end{subequations} In the equations (\ref {IV1}) and (\ref {IV2}) $e_i$s are the basis vectors for an n-dimensional vector space. It follows from (\ref {2.6}b) that ${\bf A}^{\dagger}=-{\bf A}$ for real valued $\lambda$ and thus we have \begin{equation} \partial_x({{\bf \Psi}^{(1)}}^{\dagger}{\bf \Psi}^{(2)})= 0 \label{IV3} \end{equation} for any pair of solutions of equation (\ref {2.1}a), ${\bf \Psi}^{(1)}$ and ${\bf \Psi}^{(2)}$, having the same eigenvalue. It is straightforward to show from (\ref{IV1}) and (\ref{IV2}) that \begin{equation} \Phi^{(i)\dagger}\Phi^{(j)} = \Psi^{(i)\dagger}\Psi^{(j)}=\delta_{ij} \label{IV4} \end{equation} for $i,j =1,2,.........n$. Since the set of Jost functions ${\bf \Psi_i}$ are linearly independent and the maximum number of independent Jost functions is $n$, we may express the set of Jost functions $\Phi_i$ as a linear combination of $\Psi_i$ as \begin{equation} \Phi^{(i)}(x,\lambda) = \sum_{j=1}^n \alpha_{ij}(\lambda)\Psi^{(j)}(x,\lambda) \label{IV5} \end{equation} where $\alpha_{ij}(\lambda)$ is the $(ij)$th element of scattering data matrix, which can be expressed by using (\ref{IV4}) and (\ref {IV5}) in the form \begin{equation} \alpha_{ij}(\lambda)= \Psi^{(j)\dagger}(x,\lambda)\Phi^{(i)}(x,\lambda) \label{IV6} \end{equation} The orthogonality property of the scattering data matrix elements for real eigenvaule $\lambda$, subsequently follows from (\ref{IV1}), (\ref{IV2}) and (\ref{IV6}) and thus we obtain \begin{equation} \sum_{k=1}^{n}\alpha_{ik}(\lambda)\alpha_{jk}(\lambda) =\delta_{ij} \label{IV7} \end{equation} which finally gives \begin{equation} \Psi^{(i)}(x,\lambda) = \sum_{j=1}^n \alpha^*_{ji}(\lambda)\Phi^{(j)}(x,\lambda) \label{IV8} \end{equation} It is further interesting to see, by exploiting the properties of $[\alpha_{ij}]$ and $[\alpha^*_{ij}]$ matrices, that we can write the element $\alpha^*_{ij}$ as the cofactor of the elements of the matrix $[\alpha_{ij}]$. In particular, $\alpha^*_{ni}$ element can be written as \begin{equation} \alpha_{ni}^* = (-1)^{n+i}det[\tilde{\alpha}_{ni}] \label{IV9} \end{equation} where $[\tilde{\alpha}_{ni}]$ is a $(n-1)\times (n-1)$ matrix, constructed from the $n\times n$ scattering matrix, $[\alpha_{ij}]$ with nth row and ith column being omitted, {\it i.e.} $det[\tilde{\alpha}_{ni}]$ is the minor of $\alpha_{ni}$ element of scattering matrix, $[\alpha_{ij}]$. Now by using (\ref{IV5}) and (\ref{IV9}), we obtain the following useful relations among the Jost functions $\Phi^{(i)}$ and $\Psi^{(i)}$. The first $n-1$ Jost functions in (\ref {IV5}) satisfy \begin{subequations} \begin{equation} \frac{1}{\alpha^*_{nn}(\lambda)}\sum_{j=1}^{n-1}(Adj [\tilde{\alpha}_{nn}])_ {kj}\Phi^{(j)}e^{i\lambda x} = \Psi^{(k)}e^{i\lambda x} - \frac{\alpha^*_{nk}}{\alpha^*_{nn}}\Psi^{(n)}e^{i\lambda x} \label{IV10a} \end{equation} with $k=1,........n-1$, but the $n$th. Jost function,$\Phi^{(n)}$ obey the relation \begin{equation} \frac{1}{\alpha_{nn}} \Phi^{(n)}e^{-i\lambda x} = \Psi^{(n)}e^{-i\lambda x} + \frac{1}{\alpha_{nn}} \sum_{j=1}^{n-1} \alpha_{nj}\Psi^{(j)}e^{-i \lambda x} \label{IV10b} \end{equation} \label{IV10} \end{subequations} Notice that in deriving (\ref{IV10}), we have used the following properties of the scattering data matrix. \[ \alpha^*_{nk}\delta_{ij} = \sum_{l=1}^{n-1}[\tilde{\alpha_{nk}}]_{il} (Adj [\tilde{\alpha}_{nk}])_{lj} \] It is important to mention that analyticity properties of the Jost functions and the elements of scattering matrix may be obtained from (\ref{IV10}). We are now going to derive the Gelfend, Levitan and Marchencho equation for an $n$ dimensional Lax pair. Let us consider an integral representation of the Jost function $\Psi^{(i)}$ for $(i=1,2,.....,n)$. For the first $n-1$ Jost functions we may choose the following integral representations \begin{subequations} \begin{equation} \Psi^{(j)}(x,\lambda) =e_j e^{-i\lambda x} + \int_x^{\infty}dy {\bf K}^{(j)}(x,y)e^{-i\lambda y} \label{IV11a} \end{equation} with $j=1,2,.....n-1$, while the $n$th Jost function may be written as \begin{equation} \Psi^{(n)}(x,\lambda) =e_n e^{i\lambda x} + \int_x^{\infty}dy {\bf K}^{(n)}(x,y)e^{i\lambda y}. \label{IV11b} \end{equation} \label{IV11} \end{subequations} where, $e_i,\quad i=1,2,\cdots,n$ are basis vectors for an $n$ dimensional vector space and the kernels ${\bf K}^{(j)}$ and ${\bf K}^{(n)}$ are $n$ dimensional column vectors, which may be written explicitly in the component form as \begin{subequations} \begin{equation} {\bf K}^{(j)}(x,y)=\sum_{m=1}^n K_m^{(j)}(x,y)e_m \label{IV12a} \end{equation} \begin{equation} {\bf K}^{(n)}(x,y)=\sum_{m=1}^n K_m^{(n)}(x,y)e_m \label{IV12b} \end{equation} \label{IV12} \end{subequations} \newpage Substituting (\ref{IV11}) in (\ref{IV10b}), we obtain \begin{eqnarray} \frac{1}{\alpha_{nn}}\Phi^{(n)} &=& e_n e^{i\lambda x} +\int_x^{\infty} dy {\bf K}^{(n)}(x,y)e^{i\lambda y} + \sum_{j=1}^{n-1} \frac{\alpha_{nj}}{\alpha_{nn}}e_j e^{-i \lambda x} \nonumber \\ &+& \sum_{j=1}^{n-1}\frac{\alpha_{nj}}{\alpha_{nn}}\int_x^{\infty} dy {\bf K}^{(j)}(x,y)e^{-i\lambda y} \label{IV13} \end{eqnarray} Multiplying now both sides of (34) by $\frac{1}{2\pi} \int_{-\infty}^{\infty}d\lambda e^{-i\lambda z}$, with an assumption $z>x$ and using the analyticity property of the Jost function, it follows that \begin{subequations} \begin{eqnarray} &~&\frac{1}{2\pi}\int_{-\infty}^{\infty} d\lambda \frac{\Phi^{(n)}} {\alpha_{nn}(\lambda)} e^{-i\lambda z} = {\bf K}^{(n)}(x,z) + \int_{-\infty}^{\infty}\frac {d\lambda}{2\pi}\sum_{j=1}^{n-1}\frac{\alpha_{nj}(\lambda)} {\alpha_{nn}(\lambda)}e_j e^{-i\lambda(x+z)} \nonumber \\ &+&\int_{-\infty}^{\infty}\frac{d\lambda}{2\pi}\int_{x}^{\infty}dy \sum_{j=1}^{n-1} \frac{\alpha_{nj}(\lambda)}{\alpha_{nn}(\lambda)}{\bf K}^{(j)}(x,y) e^{-i\lambda(y+z)} \\ \nonumber &~& \label{IV14a} \end{eqnarray} The L.H.S. of (\ref{IV14}a) can be simplified further by taking into account that $\frac{1}{\alpha_{nn}(\lambda)}$ is analytic in the lower half plane except at the points, say $\lambda_j^*$ with $Im \lambda_j^*>0$ and $j=1,2........N$, where $\frac{1}{\alpha_{nn}(\lambda)}$ has $N$ simple poles. Moreover, we assume that at the simple poles $\lambda_j^*$, $\Phi^{(n)}$ to be of the form \begin{equation} \Phi^{(n)}(x,\lambda_j^*)=\sum_{p=1}^{n-1}C^{(j)}_{np}\Psi^{(p)} (x,\lambda_j^*) \label{IV14b} \end{equation} With these assumptions, L.H.S. of (\ref{IV14}a) becomes \begin{equation} -i\sum_{j=1}^N\frac{e^{-i\lambda_j^*z}}{\alpha^{\prime}_{nn}(\lambda^*_j)} \sum_{p=1}^{n-1}C_{np}^{(j)}\Psi^{(p)}(x,\lambda_j^*) \label{IV14c} \end{equation} where $\alpha_{nn}^{\prime}$ denotes derivative with respect to $\lambda $. By Substituting the integral representations of $\Psi^{(p)}(x,\lambda_j^*)$ from (\ref {IV11a}), (\ref {IV14c}), {\it i.e.} the L.H.S. of (\ref {IV14}a) finally reduces to \begin{equation} -i\sum_{j=1}^N \frac{e^{-i\lambda_j^*z}}{\alpha_{nn}^{\prime}(\lambda_j^*)} \sum_{p=1}^{n-1}C_{np}^{(j)} [e_p e^{-i\lambda_j^* x} +\int_x^{\infty}dy {\bf K}^{(p)}(x,y)e^{-i\lambda_j^* y}] \label{IV14d} \end{equation} \label{IV14} \end{subequations} Let us now introduce a funtion $F_p(x+y)$ as \begin{equation} F_p(x+y)=i\sum_{j=1}^{N}\frac{C_{np}^{(j)}e^{-i\lambda^*_j(x+y)}} {\alpha_{nn}^{\prime}(\lambda^*_j)} + \int_{-\infty}^{\infty}\frac{d\lambda}{2\pi}\frac{\alpha_{np}(\lambda)} {\alpha_{nn}(\lambda)}e^{-i\lambda(x+y)} \label{IV15} \end{equation} In terms of the function, $F_p(x+y)$ in (\ref {IV15}), (\ref{IV14}a) and (\ref{IV14d}) together may be written in a compact form as \begin{subequations} \begin{equation} {\bf K}^{(n)}(x,z) + \sum_{p=1}^{n-1}e_pF_p(x+z) + \sum_{p=1}^{n-1} \int_x^{\infty}{\bf K}^{(p)}(x,y)F_p(z+y) =0 \label{IV16a} \end{equation} The inetgral equation (\ref {IV16a}) is one of the desired Gelfand Levitan Marchenko equations for the kernel ${\bf K}^{(n)}$. Other integral equations for the kernels ${\bf K}^{(p)}$ for $p=1,2,\cdots, n-1$ may be obtained from (\ref{IV10a}) and (\ref{IV11}) in a similar way like that of (\ref {IV16a}). The integral equations for the kernels ${\bf K}^{(p)}$ thus turn out to be of the form \begin{equation} {\bf K}^{(p)}(x,z) - e_n F_p^*(x+z) - \int _x^{\infty}dy {\bf K}^{(n)}(x,z)F_p^*(z+y) =0 \label{IV16b} \end{equation} \label{IV16} \end{subequations} provided $z>x$. In deriving (\ref {IV16b}) we have used the following identity \[ C_{np}^* = \alpha_{np}^*(\lambda_j)= \sum_{i=1}^{n-1}[{\tilde \alpha_{np}}(\lambda_j)]_{ki} (Adj[{\tilde \alpha_{np}}(\lambda_j)])_{il} \delta_{kl}. \] The set of coupled equations (\ref {IV16}) may be called as generalized Gelfand Levitan Marchenko equations. Substituting now (\ref {IV12b}) and (\ref{IV16b}) in (\ref{IV16a}), to a first approximation, we find the Gelfand Levitan Marchenko equation for the $p$th. component of ${\bf K}^{(n)}$: \begin{equation} K^{(n)}_p(x,z) +F_p(x+z) +\sum_{m=1}^{n-1} \int_x^{\infty}ds K^{(n)}_p(x,s) \int_x^{\infty}dyF_m(y+z)F^*_m(y+s) =0 \label{IV17} \end{equation} which will be used later to find the soliton solutions for VHNLS equataion. \section{\bf N Soliton Solutions} To obtain soliton solutions, let us associate dynamical fields $q_i(x,t)$ with the kernels, $K^{(n)}_i$ in (\ref {IV17}). Substituting (\ref{IV11b}) into Lax equation (\ref{2.1}a) we find \begin{equation} q_i(x)= -2K^{(n)}_i (x,x) \label{V1} \end{equation} and consequently, it is evident that the solution of (\ref {IV17}) gives rise to soliton solutions in terms of scattering data elements. But before going to solve (\ref {IV17}), we first compute time evolution of scattering data element. Notice that as $\vert x \vert \rightarrow \infty$, the Lax equation (\ref {2.2b},\ref {2.3}b) leads to \begin{equation} \frac{\partial \Psi}{\partial t} =-4i\epsilon \lambda^3\Sigma \label{V2} \end{equation} which, in turn, determines time evolution of the scattering data elements. For example, scattering data elements, $\alpha_{nj}(\lambda,t)$ for $j=1,2, \cdots, n-1$ evolve with time as \begin{subequations} \begin{equation} \alpha_{nj}(\lambda,t) = \alpha_{nj}(\lambda,0)e^{-8i\epsilon \lambda^3t} \label{V3a} \end{equation} while the element $\alpha_{nn}(\lambda,t)$ is time invariant: \begin{equation} \alpha_{nn}(\lambda,t) = \alpha_{nn}(\lambda,0), \label{V3b} \end{equation} \label{V3} \end{subequations} which eventually justifies our conjecture in section 3 that $\alpha_{nn}$, indeed, can be associated with the conserved quantities. From (\ref {IV8}) and (\ref {IV14b}) it follows that the coefficients $C^{(j)}_{np}$ also satisfy the similar time dependence as $\alpha_{nj}$ in (\ref {V3a}) and thus \begin{equation} C^{(j)}_{np}(\lambda,t) = C_{np}^{(j)}(\lambda,0)e^{-8i\epsilon \lambda^3 t} \label{V4} \end{equation} If we now restrict ourselves to soliton sector, $\alpha_{np}(\lambda)$ becomes trivial and the function $F_p(x+y)$, in this case, reduces to \[ F_p(x+y)=i\sum_{j=1}^{N}\frac{C_{np}^{(j)}e^{-i\lambda^*_j(x+y)}} {\alpha_{nn}^{\prime}(\lambda^*_j)}. \] Time dependence of the function $F_p(x+y)$ may be obtained immediately from (\ref {V4}) as \begin{equation} F_p(x+y)=i\sum_{j=1}^{N}\frac{C_{np}^{(j)}(\lambda,0)e^{-8i\epsilon \lambda^{*3}_jt} e^{-i\lambda^*_j(x+y)}}{\alpha_{nn}^{\prime}(\lambda^*_j)}. \label{V6} \end{equation} To solve the integral equation (\ref {IV17}) for the soliton solutions, the kernels $K^{(n)}_p(x+y)$ are assumed to be of the form \begin{equation} K_p^{(n)}(x+y) = \sum_{j=1}^{N}\omega_{pj}(x,t)e^{-i\lambda^*_j y} \label{V7} \end{equation} Substituting (\ref {V6}) and (\ref {V7}) in (\ref {IV17}), we obtain a set of $n-1$ algebraic equations: \begin{eqnarray} K^{(n)}_p(x,x)&+&F_p(x+x)-\sum_{j,k,l=1}^{N}\sum_{r=1}^{n-1}\frac{\omega_{pl} e^{-i\lambda_j^*x}}{(\lambda_k-\lambda_j^*)(\lambda_k-\lambda_l^*)}\cdot \nonumber\\ &~&\frac{C^{(j)}_{nr}(0)C^{*(k)}_{nr}(0)}{\alpha^{\prime}_{nn}(\lambda_j^*) \alpha^{*\prime}_{nn}(\lambda_k)} e^{i(2\lambda_k-\lambda_j^*-\lambda_l^*)x+8i\epsilon (\lambda_k^3-\lambda_j^{*3})t}=0 \label{V8} \end{eqnarray} Solving (\ref {V8}) for $K_p(x,x)$ and subsequently using (\ref {V1}), the $N$ soliton solutions for each dynamical field $q_i(x,t)$ may be expressed as \begin{equation} q_i(x,t) = -2\sum^{N}_{j=1}({\bf B}{\bf C}^{-1})_{ij}e^{-i\lambda_j^*x} \label{V9} \end{equation} where, ${\bf B}$ and ${\bf C}$ are respectively $(n-1)\times N$ and $N\times N$ matrices whose explicit forms are given by \[ ({\bf B})_{ij} = i C^{(j)}_{ni}(0)e^{-8i\epsilon\lambda_j^{*3}t - i \lambda_j^{*}x} \] \[ ({\bf C})_{ij} = \sum_{p=1}^{n-1}\sum_{k=1}^N \frac{C^{(j)}_{np}(0) C^{(k)}_{np}(0)e^{-i(\lambda_i^*+\lambda_j^*-2\lambda_k )x + 8i\epsilon(\lambda_k^3-\lambda_j^{*3})t}}{\alpha^{\prime}_{nn}(\lambda_j^*) \alpha^{*\prime}_{nn}(\lambda_k)(\lambda_k-\lambda_j^*)(\lambda_k-\lambda_i^*)} -\delta_{ij} \] Let us now consider an explicit form of one soliton. One soliton solution, which follow from (\ref {V9}), may be written in the following form \begin{equation} q_i(x,t) = \frac{2F_i(x+x)}{1 + \sum_{p=1}^{n-1} \vert F_p(x+x) \vert^2 \frac{1}{(\lambda_1 -\lambda_1^*)^2}} \label{V10} \end{equation} where \[ F_i(x+y) = i\frac{C_{ni}^{(1)}(\lambda;0)} {\alpha_{nn}^{\prime}(\lambda^*_1)}e^{-8i\epsilon \lambda^{*3}_{1}t - i \lambda^*_1(x+y)}. \] $q_i(x,t)$ in (\ref {V10}) may be expressed in a more conventional way, by choosing the position of the pole at $\lambda_1=\frac{1}{2}(-\xi+i\eta)$ and by introducing $e^{\gamma_i+i\delta_i}=\frac{C_{ni}^{(1)}(\lambda,0)} {2\eta \alpha_{nn}^{\prime}(\lambda^*_1)}$ and thus \begin{equation} q_i(x,t) = 4i\eta \frac{e^{\gamma_i-\eta x+\epsilon\eta(\eta^2-3\xi^2)t +i(\delta_i+\xi x+\epsilon\xi^3t-3\epsilon\xi^2\eta t)}}{1+\sum_{p=1}^{n-1} e^{2\gamma_p}e^{-2\eta x+2\epsilon(\eta^3-3\eta\xi^2)t}} \label{V11} \end{equation} which can further be simplified as \begin{equation} q_i(x,t) = \frac{2i\eta e^{\gamma_i-\Gamma_n + iQ_i}}{coshP} \label{V11a} \end{equation} by introducing \begin{eqnarray*} e^{2\Gamma_n} &=& \sum_{p=1}^{n-1}e^{2\gamma_p}\\ P(x,t) &=& \eta x-\epsilon \eta(\eta^2-3\xi^2)t-\Gamma_n\\ Q_i(x,t)&=&\xi x+\epsilon \xi(\xi^2-3\xi\eta)t+\delta_i\\ \end{eqnarray*} Once again, if we specialize to HNLS equation, each independent field $q_i(x,t)$, depending on the models, reduces either to $q$ or to $q^*$ and as a consequence $q_i$ in (\ref {V11}) yields to \begin{equation} q(x,t) = \frac{2i\eta e^{i{\tilde B}}}{\sqrt{(n-1)} cosh{\tilde A}} \label{V12} \end{equation} where, ${\tilde A}= \eta x-\epsilon\eta(\eta^2-3\xi^2)t-\gamma -\frac{1}{2}ln(n-1)$ and ${\tilde B}=\xi x+\epsilon\xi(\xi^2-3\xi\eta)t+\delta$. It is interesting to note that we have obtained precisely the same expression for one soliton in \cite {11}. \vspace{.5 cm} \section{\bf Generalized Landau Lifshitz type equation as the Gauge equivalence system} We now show an interesting connection between the VHNLS equation and the generalized Landau Lifshitz type equation by exploiting the gauge equivalence of the Lax pairs of these two dynamical systems. The procedure is similar to that between the nonlinear Schr{\" o}dinger equation and the standard Landau Lifshitz equation \cite {14}. Under a local gauge transformation, the Jost function, ${\bf \Psi}(x,t,\lambda)$ changes to \begin{equation} \tilde{{\bf \Psi}} = g^{-1}(x,t){\bf \Psi}(x,t,\lambda) \label{III1} \end{equation} where $g(x,t) = {\bf \Psi}(x,t,\lambda)\vert_{\lambda =0}$. We claim that $g(x,t)$ is an element of $SU(n)$ group. As a consequence of the gauge transformation (\ref {III1}), the Lax equations (\ref {2.2},\ref{2.3}) become \begin{subequations} \begin{eqnarray} \tilde{{\bf \Psi_x}} &=& \tilde{{\bf U}}(x,t,\lambda)\tilde{{\bf \Psi}}\\ \tilde{{\bf \Psi_t}} &=& \tilde{{\bf V}}(x,t,\lambda)\tilde{{\bf \Psi}} \end{eqnarray} \label{III2} \end{subequations} where $\tilde{U}$ and $\tilde{V}$are the new gauge transformed Lax pair, given by \begin{subequations} \begin{eqnarray} \tilde{{\bf U}}(x,t,\lambda) &=& g^{-1}({\bf U}-{\bf U}_0)g\\ \tilde{{\bf V}}(x,t,\lambda) &=& g^{-1}({\bf V}-{\bf V}_0)g \end{eqnarray} \label{III3} \end{subequations} with ${\bf U}_0 ={\bf U}\vert_{\lambda=0}=g_x(x,t)g^{-1}(x,t)$ and ${\bf V}_0 ={\bf V}\vert_{\lambda=0}=g_t(x,t)g^{-1}(x,t)$. This leads to \begin{equation} {\bf A}=g_x(x,t)g^{-1} \nonumber \end{equation} Since ${\bf A}$ belongs to $su(n)$ algebra, $g(x,t)$ obviously belongs to $SU(n)$ group, which justifies our claim. We may now identify the spin field of the Landau Lifshitz type equation as \begin{equation} S =g^{-1}(x,t){\bf \Sigma}g(x,t), \quad\quad\quad S^2 =1 \label{III4} \end{equation} With this identification, the gauge transformed Lax pair (\ref{III1}) may be expressed in terms of the spin field $S$ (\ref{III4}) and its derivatives only, yielding \begin{subequations} \begin{eqnarray} \tilde{{\bf U}} &=&-i\lambda S\\ \tilde{{\bf V}} &=& -4i\epsilon \lambda^3S + 2\epsilon\lambda^2SS_x +i\epsilon\lambda(S_{xx} + \frac{3}{2}SS_x^2) \end{eqnarray} \label{III5} \end{subequations} In deriving (\ref {III5}) we have used the following important identities \begin{eqnarray*} SS_x &=& 2g^{-1}{\bf A}g\\ SS_x^2 &=& -4g^{-1}{\bf \Sigma}{\bf A}^2g \\ S_{xx} &+& SS_x^2 = 2 g^{-1}{\bf \Sigma}{\bf A}_xg \end{eqnarray*} The zero curvature condition of (\ref {III5}), namely \[ \tilde{{\bf U}}_t - \tilde{{\bf V}}_x + [\tilde{{\bf U}}, \tilde{{\bf V}}]=0 \] ultimately leads to the generalized Landau Lifshitz type equation \begin{equation} S_t+\epsilon S_{xxx}+\frac{3}{2}\epsilon (S_x^3+SS_{xx}S_x+SS_xS_{xx}) = 0 \label{III6} \end{equation} with$S\in SU(n)/(U(n-1))^n$. \vspace{.5 cm} \section{\bf Conclusion} We have formulated a generalized IST method for an $n$ dimensional Lax oparator. Subsequently, the generalized IST method is used to obtain $N$ soliton solutions for a multi-component generalization of HNLS equation, proposed by us. We have, although, obtained soliton solutions for VHNLS equation, IST method developed here is quite general and is applicable for obtaining soliton solutions for all integrable nonlinear equations. The $sech$ structure of one soliton solution of VHNLS equation is quite interesting, particularly in the context of nonlinear optics, since it can easily be produced from the output of a mode locked laser. We have shown the integrability of VHNLS equation also in the Liouville sense. This has been acheived first by finding a set of coupled Riccati equations and subsequently by identifying the generating function for the conserved charges. It is found that the last diagonal element of the scattering data matrix may be identified as the generating function of the conserved charges. This is also confirmed from the time evolution of scattering data matrix elements. We have also established an intriguing relationship between the VHNLS equation and a generalized Landau Lifshitz equation, where the spin field $S\in SU(n)/(U(n-1))^n$. Moreover, we have obtained a Lax pair for the spin system implying direct integrability of the generalised Landau Lifsitz equation. \vspace{1.0 cm} S.N. would like to thank CSIR, Govt. of India for financial support and for the award of Junior Research Fellowship. Electronic address : <EMAIL>, $^{\dag}[email protected]
\section{Introduction} The two most powerful FR~II radio sources in the nearby Universe -- Cyg~A and 3C295 -- are each located at the centre of a dense, moderately rich cluster of galaxies. While such an environment is exceptional for a low-redshift FR~II galaxy, it appears to be common around powerful radio objects at earlier epochs. Above a redshift of 0.5, radio-loud objects (both the quasars and radio galaxies) are inferred to lie in clusters of galaxies of moderate optical richness. The evidence for such an environment includes optical and near-IR galaxy counts (Yee \& Green 1987; Yates et al 1989; Hill \& Lilly 1991; Ellingson et al 1991; Dickinson 1997), high gas pressures within a radius of 30{\rm\thinspace kpc}\ (Crawford \& Fabian 1989; Forbes et al 1990; Bremer et al 1992; Durret et al 1994), cD-type host galaxy profiles (Best et al 1998), a gravitational arc (Deltorn et al 1997), and lensing shear of surrounding field galaxies (Bower \& Smail 1997). The properties of the radio source itself also imply the presence of a confining medium: a large-scale working surface on which the jets form the radio lobes; a steep radio spectrum; and a high minimum pressure in regions of relaxed radio structure (Bremer et al 1992). A Faraday depolarization asymmetry (Garrington \& Conway 1991), the distortion and compression of high-redshift radio source morphologies (Hintzen et al 1983; Barthel \& Miley 1988) and sources with very high Faraday rotation measures (Carilli et al 1994; Carilli et al 1997) all corroborate the inference of a dense, clumpy medium surrounding the radio source. Thus it appears that the deepest potential wells we can readily pinpoint at $z\ge1$ are those around powerful radio sources. The cluster distribution at high redshift can provide a stringent cosmological test (see e.g. Donahue et al 1998), and can also be compared to the X-ray luminosity function of clusters at low redshift (eg Ebeling et al 1997). Whilst it may result in a sample of clusters biased to only those that can host an active nucleus, using radio sources to identify the location of deep potential wells is a promising way of finding clusters out to and beyond a redshift $z\sim1$ (Crawford 1997). Current X-ray surveys of clusters detected from the ROSAT All-Sky Survey (eg Ebeling et al 1998) do not reach sufficiently faint flux levels, and studies of deep serendipitous X-ray pointings (e.g. Rosati et al 1998) cover only a small fraction of the sky. The first step, however, is simply to confirm that powerful radio quasars beyond a redshift of a half really do lie at the centre of clusters of galaxies. The clearest way to determine directly the presence of a cluster of galaxies is to detect thermal X-ray emission from its hot intracluster medium. A certain degree of success has been achieved in detecting and spatially resolving the X-ray emission around distant ($0.5<z<2$) radio galaxies using ROSAT (Crawford \& Fabian 1993; Worrall et al.\ 1994; Crawford \& Fabian 1995a, 1996a,b; Crawford 1997; Dickinson 1997; Hardcastle, Lawrence \& Worrall 1998; Carilli et al 1998). Any X-rays emitted by the central bright nucleus of radio galaxies are assumed to be absorbed along the line of sight, as observed for the powerful low-redshift radio galaxy Cygnus-A (Ueno et al 1994). The inferred bolometric luminosity of the X-ray sources associated with the radio galaxies is $\sim0.7-18\times10^{44}$\hbox{$\erg\s^{-1}\,$}, easily compatible with that expected from moderately rich clusters of galaxies around the radio sources. There could also be a contribution to the extended X-ray emission from inverse Compton scattering of the hidden quasar radiation (eg Brunetti, Setti \& Comastri 1997). In the case of radio quasars, however, the X-ray detection of the spatially extended environment is complicated by the presence of bright spatially-unresolved X-ray emission from the active nucleus. The ROSAT PSPC did not combine the necessary sensitivity with a sufficiently good point-response function, needed to both detect and resolve any cluster emission around quasars. Upper limits of $1.6-3.5\times10^{44}$\hbox{$\erg\s^{-1}\,$} (in the 0.1-2.4{\rm\thinspace keV}\ rest-frame band) to X-ray emission from the environment of three radio-loud quasars have been derived from ROSAT HRI data (Hall et al 1995, 1997) assuming the cluster emission profile is modelled by a King law. We have also obtained ROSAT HRI data to spatially resolve and detect the extended emission from the intracluster medium around each of a small sample of intermediate-redshift radio-loud quasars. The detection of such a component is, however, complicated by the wobble of the spacecraft during the observation. This occurs on a $\sim402$~s period, and when the attitude of the spacecraft is not well reconstructed, leads to smearing of the point-spread function (PSF). The bright emission from the quasar nucleus can then contaminate the outer regions where we hope to detect emission from any surrounding cluster, and this has so far hindered our progress in interpreting the data. In this paper, however, we present an analysis of our ROSAT HRI data taken of seven intermediate-redshift (0.1$<z<$0.8) radio-loud quasars, which employs a new correction for the spacecraft wobble derived by Harris et al. (1998). A contemporaneous and independent analysis of an overlapping dataset using this technique has been carried out by Hardcastle \& Worrall (1999), who obtain similar results. \section{Observations and analysis} We use the ROSAT data of intermediate-redshift, radio-loud quasars for which there is prior evidence from other wavebands for a cluster environment (see notes on individual quasars for details). We preferentially selected quasars of only moderate X-ray luminosity in order to minimise the contrast between the nuclear emission and any cluster emission. These targets were supplemented by data available from the ROSAT public archive on 3C273 and 3C215. We also include the observations of H1821+643 to form a comparison to the results of Hall etal (1997). The observations used, and details of the quasars are listed in Table~\ref{tab:obslog}. \onecolumn \begin{table} \caption{Target sample and results \label{tab:obslog}} \begin{tabular}{lccccccc} & & & & & & & \\ & & & & & & & \\ Quasar& RA & DEC & Redshift & N$_H$ & ROR & Exposure & Roll angle \\ & (J2000) & (J2000) & $z$ & ($10^{20}$\hbox{$\cm^{-2}\,$}) & & (sec) & interval \\ 3C48 & 01 37 41.3 & 33 09 35 & 0.367 & 4.54 & 800634n00 & 37362 & 1-60000 \\% Bremer 3C215 & 09 06 31.9 & 16 46 13 & 0.412 & 3.65 & 800753a01 & 39173 & 27000-58000 \\% Hall & & & & & 800753n00 & 17231 & 14000-22000 \\ & & & & & 800718n00 & 16148 & 1-50000 \\ 3C254 & 11 14 38.5 & 40 37 20 & 0.734 & 1.90 & 800721n00 & 29162 & 29000-49000 \\ 3C273 & 12 29 06.7 & 02 03 09 & 0.158 & 1.79 & 701576n00 & 68154 & 1-32000 \\ 3C275.1& 12 43 57.7 & 16 22 53 & 0.555 & 1.99 & 800719n00 & 25396 & 9000-32000 \\% Bremer 3C281 & 13 07 53.9 & 06 42 13 & 0.602 & 2.21 & 800635a01 & 20299 & 12000-24000 \\% Bremer & & & & & 800635n00 & 18220 & 1-14000 \\ 3C334 & 16 20 21.5 & 17 36 29 & 0.555 & 4.24 & 800720n00 & 28183 & 17000-42000 \\% Bremer H1821+643 & 18 21 57.3 & 64 20 36 & 0.297 & 4.04 & 800754n00 & 29427 & 23000-31000 \\%Hall\\%RQQ/galaxy & & & & & & & \\ \end{tabular} \\ Notes:\\ Further possible observations of 3C215 (800718a01), 3C194 (800803n00) and 3C280 (800802n00) all had too few photons in the source for a satisfactory wobble correction. \\ The hydrogen column density along the line of sight to each quasar (N$_H$) is calculated from the data of Stark et al (1992). \\ ROR in column 6 is the ROSAT observation request sequence number. \\ The roll angle interval tabulated in the final column is given in sequence numbers taken from the attitude file, which contains the attitude information of the telescope (such that roll angle x1 at time y1 gives sequence number 1). The time interval of the spacecraft clock is 1{\rm\thinspace s}. Sequence numbers from 1 to 34000 thus mean that the data is extracted from between the time y1 and the time y34000 (34000{\rm\thinspace s}). \\ \end{table} \twocolumn \subsection{Wobble-correction technique} The spatial analysis of ROSAT HRI observations is often complicated by smearing of the image on the order of 10 arcsec (Morse 1994). This degradation of the instrinsic resolution of the HRI instrument (5 arcsec) can be induced by errors in the aspect solution associated with the wobble of the ROSAT spacecraft, or with the reacquisition of the guide stars. To counteract this effect, we use the wobble-correction technique of Harris et al (1998) which minimizes the spatial smearing of the sources. The technique is based on the simple assumption that in the case of a stable roll angle (i.e. the same guide star configuration) the aspect error is repeated through each cycle of the wobble. We thus select data only from the longest constant roll angle interval of an observation and folded these data over the ROSAT 402~s wobble phase. This phase was grouped into a number of intervals (5, 10 or 20) in order to calculate phase-resolved sub-images. The centroid of each of the sub-images was calculated by fitting a 2d-gaussian to the brightest source near the field center. The wobble-corrected HRI events list is reconstructed by adding the sub-images which have been shifted to the centroid position of the uncorrected HRI image. If a sub-image contains too few photons with which to determine the centroid, then the position of the sub-image is not changed. This makes sure that source extension cannot be produced by a failure of the centroid determination in one or more sub-images. For X-ray sources with countrates $\sim$0.1\hbox{$\ct\s^{-1}\,$} the method can reduce the full width half maximum by about 30 per cent (cf. Harris et al. 1998). The efficacy of this technique has been tested on HRI observations of low-luminosity X-ray AGN, some of which show no sign of extended X-ray emission once corrected for wobble (Lehmann et al. 1999). We cannot apply the wobble-correction technique to all possible data on suitable targets in the literature, as it can only be applied to observations that have a sufficient number of photons within a constant spacecraft roll angle (cf. Table~\ref{tab:obslog}). \subsection{Radial profile analysis} Our aim is to search for spatially extended X-ray emission originating in any intracluster medium around the quasars. We have derived a background-subtracted radial profile from the un-corrected and wobble-corrected HRI data of the X-ray source associated with each quasar. First we determined the centroid position of the X-ray source by fitting a 2D Gaussian. Then we calculated the counts per arcsec$^2$ within the rings around the centroid position from 0 arcsec to 2.5 arcsec in steps of 0.5 arcsec, from 2.5 arcsec to 15 arcsec in steps of 2.5 arcsec, from 15 arcsec to 100 arcsec in steps of 12.5 arcsec and from 100 arcsec to 1000 arcsec in steps of 100 arcsec. The background value was calculated as the median from the 6 rings between 300 arcsec and 900 arcsec. Finally we subtracted the background value from each ring. First, however, we need to accurately model the HRI PSF which will allow us to remove the contaminating spill-over of light from the bright quasar nucleus. The standard HRI PSF of David et al (1995) does not provide a good fit to the radial profile of observed point sources (see Fig~\ref{fig:emppsf}). This deviation is particularly acute where the PSF shows a sharp drop at radii between 10 and 30 arcsec, where we expect the contrast between the nuclear source and any extended cluster emission to begin to show. Instead we have determined a good analytical characterization to an empirical PSF derived from observations of 21 ROSAT Bright survey stars, each of which has undergone the same wobble correction procedure as the quasar data. Our best fit is the sum of two Gaussians and a power-law component (Fig~\ref{fig:emppsf}). Assuming that this PSF forms a good model for the spillover of quasar nuclear light, we fix the relative normalizations and widths of the three components and leave only the overall normalization $n$ of this profile as the only free variable in the function: $$ I(r) = n \{ e^{-0.5(r/4.5001)^2} + 4.376 e^{-0.5(r/2.8644)^2} + 0.9346 r^{-1.6569} \}$$ \begin{figure} \psfig{figure=emppsf.ps,width=0.45\textwidth,angle=270} \caption{ \label{fig:emppsf} The best-fit analytic model (dashed line) to the empirically-derived HRI PSF (circle markers), shown with an arbitrary flux scaling. The solid line shows the standard HRI PSF of David et al (1995) for comparison. } \end{figure} We follow the same procedure for the profile fitting of each quasar. First we fit the profile by the empirically-derived HRI PSF alone, allowing its normalization to vary freely. We then fit this PSF (with normalization still free to vary) in combination with each of two models chosen to represent any extended emission. The first model is a broken power-law, with slope $r^{-1}$ for radii $r<R$ and $r^{-2.1}$ for $r\ge R$, where the break $R$ and the absolute normalization of the extended component are free to vary. This is an approximation to the X-ray surface brightness profile of the gas in a typical cluster of galaxies containing a cooling flow, and $R$ then corresponds to the cooling radius (eg Crawford \& Fabian 1995b). The second model employed is a projected King law, with index fixed at -1.5, and the core radius $R$ and normalization left as free parameters. Given the errors inherent in whether such simple models truly characterize the extended emission, we do not convolve the extended emission models with the PSF. The relative normalization between the PSF and extended components are not always very well determined, so we also derive what should be regarded as a lower limit to the presence of any extended component by assuming the nuclear emission accounts for all the light in the X-ray core. We fit the PSF to the quasar radial profile within the inner 1--5 arcsec and then subtract this model and fit the residuals by each of the cluster models. We execute these 5 model fits to the profiles out to a radius of 50 arcsec (11 data points), yielding 10, 8, and 9 degrees of freedom for the psf only, psf+extended component models, and the fit of the extended component model to the residual after subtraction of the normalised HRI PSF. We then repeat the fits to the profile out to a radius of 100 arcsec (15 data points), yielding 14, 12 and 13 degrees of freedom to the fits as above. The fitting analysis is carried out first for each quasar image in the absence of any wobble correction, and then for the images corrected using different phase intervals. The detailed results are summarized in Table~2, where the reduced-$\chi^2$ is given for each fit. We tabulate the best-fit parameters of the profile fits out to a radius of 50 arcsec and then 100 arcsec in turn: $R$ (in arcsec) representing either the break in the broken power-law model, or the core radius in the King law model; the integrated luminosity from the extended component as a percentage of the total luminosity of the X-ray source; the X-ray luminosities (in the observed 0.1-2{\rm\thinspace keV}\ energy band) of the quasar component ($L_X^{QSO}$) and that of the cluster component ($L_X^{cl}$) assuming a power-law of photon index 2 and thermal bremsstrahlung emission at a temperature of $kT=4{\rm\thinspace keV}$ respectively. (At the redshift of our quasars this observed band carries about half of the bolometric luminosity for the thermal spectrum.) The errors are derived from propagating the $\Delta\chi^2=1$ confidence limits of the fit parameters. Errors are not shown when the fit was insufficiently robust to extract errors on all parameters of interest. Table~2, however, demonstrates the full range of values obtained from the ten model fits employed for each of the phase intervals and allows one to assess the variation of each parameter from the systematic uncertainties of PSF normalization and extended component model employed. A comparison of some of the better fits to the radial profile of each quasar (those shown in bold font in Table~2) are displayed in Figure~\ref{fig:profs}. These plots clearly show that there are significant differences between the PSF-only fit to the profile, and the fits that include a model for extended emission. In all this analysis we necessarily assume that any extended component is both centred on the quasar (in no case do we see any evidence for a secondary off-centre peak), and derive its properties such as scale and luminosity assuming that it is at the redshift of the quasar. The present data cannot rule out a contribution to the extended component of X-ray emission from the active nuclei of close companion galaxies to each of the quasars. Such emission would, of course, provide further support for a clustered environment. The probability of getting an unassociated X-ray source within an aperture of 1 square arcminute centred on a quasar is less than $10^{-3}$, at the flux level of the extended emission. Thus there is little chance of the extended emission component being due to contamination by fore- or back-ground sources. \section{Results for individual quasars} \subsection{3C48} Given its proximity to a very luminous source of photoionizaton, the low ionization state observed in the spatially extended oxygen line emission around this 3C48 led Fabian et al (1987) to deduce a high density environment around this quasar. The inferred gas pressure of 3-8$\times10^5$\hbox{$\cm^{-3}\K$}\ within 30{\rm\thinspace kpc}\ of the quasar core is consistent with confinement of the extended emission-line region by an intracluster medium There is, however, no strong evidence for a rich cluster of galaxies from optical images (Yee, Green \& Stockman 1986; Yates etal 1989). The fit to the radial profile is substantially improved by the addition of an extended component, the best fits being obtained in all cases when this is represented by a King law. The extended component requires a very consistent value for the core radius $R$ of around 5-6 arcsec in all fits (1 arcsec corresponds to 6.2{\rm\thinspace kpc}\ at the redshift of the quasar\footnote{We assume a value for the Hubble constant of $H_0=50$\hbox{$\kmps\Mpc^{-1}$}\ and a cosmological deceleration parameter of $q_0=0.5$ throughout this paper.}), and accounts for 10-16 per cent of the total X-ray source. The full variation of its luminosity is $5-10\times10^{44}$\hbox{$\erg\s^{-1}\,$}, with most of the values derived being to the lower end of this range. \subsection{3C215} This quasar lies in a densely clustered environment (Ellingson et al.\ 1991; Hintzen 1984), and the radio source has a very complex structure suggestive of deflection and distortion of the radio jet to the south-east by some external medium. The two sides of the radio source show asymmetric Faraday depolarization, which can be interpreted as due to differing lines of sight through a depolarizing cluster medium (Garrington, Conway \& Leahy 1991). Crawford \& Fabian (1989) inferred a high gas pressure of over $3\times10^{5}$\hbox{$\cm^{-3}\K$}\ from the ionization state of the extended line emission within 30{\rm\thinspace kpc}\ of the quasar nucleus. We have extracted radial profiles from three observations of 3C215. Smearing of the image seems to have badly affected observation number 800753n00, as $R$ decreases substantially after correction of the image. In all cases, the reduced $\chi^2$ of the fit improves from the addition of an extended component to the HRI PSF model, although there is little preference shown between the King and broken power-law models. In the fits allowing free normalization of the PSF component the $R$ derived is in the range 4-9 arcsec (where 1 arcsec corresponds to 6.5{\rm\thinspace kpc}\ at the redshift of the quasar), with the broken power-law model always yielding the larger values of $R$. The luminosity of the extended component ranges over 1.6-6.3$\times10^{44}$\hbox{$\erg\s^{-1}\,$}, and account for 11-40 per cent of the total X-ray emission from this source; the higher values are derived from the King law models. \subsection{3C254} 3C254 is a quasar with a very asymmetric radio source, and was discovered by us (Forbes et al 1990) to lie in a spectacular emission-line region extending out to radii of 80{\rm\thinspace kpc}. The optical nebula is again at low ionization, with an inferred pressure of $>10^6$\hbox{$\cm^{-3}\K$}\ within 30kpc (Forbes et al 1990; Crawford \& Vanderriest 1997). The kinematics and distribution of the line-emitting gas in combination with the radio morphology strongly suggest that the radio plasma to the east of the quasar is interacting with a dense clumpy environment (Bremer 1997; Crawford \& Vanderriest 1997). The radio source itself also shows asymmetric depolarization (Liu \& Pooley 1991). Optical continuum images show an overdensity of faint objects around the quasar consistent with a location in a compact cluster or group (Bremer 1997). The HRI image of 3C254 shows a detached secondary source of X-ray emission approximately 35 arcsec east of the quasar, but at a very low fraction of the total quasar luminosity. There is no optical counterpart to this source on the Space Telescope Science Institute Digitized Sky Survey. The fitting of the profile is improved by the addition of an extended component to the HRI PSF model, although it cannot discriminate between a King or broken power-law model. The extended emission contains 12-19 per cent of the total X-ray luminosity of the source. The characteristic scale length $R$ of the extended emission component varies over 9-15 arcsec (1 arcsec corresponds to 8.1{\rm\thinspace kpc}\ at the redshift of the quasar), with a luminosity of 5-9$\times10^{44}$\hbox{$\erg\s^{-1}\,$}. \subsection{3C273} 3C273 is famous for its jet, which can be seen in the outermost contours of our corrected images. The jet is known to emit X-rays at around 16 arcsec (nearer the core than the radio and optical features of the jet), but the emission from the jet contributes less than 0.5 per cent of the total X-ray luminosity of the source associated with the quasar (Harris \& Stern 1987). From optical images, 3C273 may be a member of a poor cluster of galaxies (Stockton 1980). All the wobble-corrected profiles of 3C273 require the addition of an extended component (preferably a King law). Each of the models for the extended emission yields slightly (and consistently) different results: the broken power-law fits tend to have $R$ of 8.7 arcsec (where 1 arcsec corresponds to 3.6{\rm\thinspace kpc}\ at the redshift of the quasar) and contain 5 per cent of the total X-ray luminosity, at 7$\times10^{44}$\hbox{$\erg\s^{-1}\,$}. The King law fits show a greater variation on the parameters, but have $R$ of only $\sim$4 arcsec, and $\sim11$ per cent of the total luminosity at 1.3-2.0$\times10^{45}$\hbox{$\erg\s^{-1}\,$}. The extended emission we find in the environs of this quasar is sufficiently luminous that it cannot easily be ascribed to the jet. \begin{figure*} \vbox{ \hbox{ \psfig{figure=3c48prof.ps,width=0.4\textwidth,angle=270} \psfig{figure=3c215prof.ps,width=0.4\textwidth,angle=270} }\hbox{ \psfig{figure=3c254prof.ps,width=0.4\textwidth,angle=270} \psfig{figure=3c273prof.ps,width=0.4\textwidth,angle=270} }\hbox{ \psfig{figure=3c275.1prof.ps,width=0.4\textwidth,angle=270} \psfig{figure=3c281prof.ps,width=0.4\textwidth,angle=270} }} \caption{ \label{fig:profs} The radial profile of the X-ray emission from each of the quasars out to 50 arcsec, with the exception of 3C275.1 where the profile is shown out to 100 arcsec. The phase interval used for this radial profile is shown in brackets after the quasar name in each plot. The data are shown as solid circle markers, with the best fit of the empirical HRI PSF on its own plotted as a dashed line. The best-fit models of the PSF+King law and the PSF+broken power-law model (upper solid and dotted lines respectively) are plotted, as well as just the extended component to each of these fits (the lower solid and dotted lines respectively). The fits shown in these plots are marked by bold font in Table~2.} \end{figure*} \twocolumn \addtocounter{figure}{-1} \begin{figure} \psfig{figure=3c334prof.ps,width=0.4\textwidth,angle=270} \psfig{figure=1821prof.ps,width=0.4\textwidth,angle=270} \caption{ } \end{figure} \subsection {3C275.1} 3C275.1 was the first quasar discovered from optical galaxy counts to be located at the centre of a rich cluster of galaxies (Hintzen et al 1981; Hintzen 1984; Ellingson et al 1991). The radio source is only slightly bent, but the two sides display an asymmetry in the Faraday depolarization (Garrington et al 1991). The quasar is embedded in a host galaxy with a continuum spatial profile and absolute magnitude typical of a bright cluster cD; this is in turn surrounded by a large (100{\rm\thinspace kpc}) optical emission-line nebula (Hintzen \& Romanishin 1986; Hintzen \& Stocke 1986). Crawford \& Fabian (1989) deduce a pressure within this gas of $>3\times10^5$\hbox{$\cm^{-3}\K$}\ at radii of $<$20{\rm\thinspace kpc}\ and thus also the presence of an intracluster medium (Crawford \& Fabian 1989). The fit to the X-ray radial profile improves slightly with the addition of an extended component to the HRI PSF. The models all imply a surprisingly broad core, so much so that fitting a second component to data only within a 50 arcsec radius is not robust. The King and broken power-law models yield very similar results, with $R\sim$36 arcsec (where 1 arcsec corresponds to 7.4{\rm\thinspace kpc}\ at the redshift of the quasar), and 24 per cent of the total X-ray luminosity of the source at $L_X\sim3\times10^{44}$\hbox{$\erg\s^{-1}\,$}. \subsection{3C281} 3C281 is known to lie in a rich cluster (Yee \& Green 1987) and Bremer et al.\ (1992) infer a pressure exceeding 2$\times10^6$\hbox{$\cm^{-3}\K$}\ in the extended emission-line gas within a radius of 20kpc. After the wobble-correction has been applied, the X-ray source associated with the quasar shows a distinct elongation to either side of the core along a position angle of 45$^\circ$ west of north. The direction of this elongation is, however, at odds with that of the radio source which has an axis 10$^\circ$ east of north. The profiles extracted from both observations of 3C281 both show an improved fit from the addition of an extended component (preferably a King law). The fractional luminosity of this component is high, ranging over 42-67 per cent at 1.1-1.7$\times10^{45}$\hbox{$\erg\s^{-1}\,$} (the higher values obtained with the King model fits). The characteristic radius of this component is 5-6 arcsec (where 1 arcsec corresponds to 7.6{\rm\thinspace kpc}\ at the redshift of the quasar). \subsection{3C334} This quasar lies in a clustered environment (Hintzen 1984) and has a pressure within the extended emission-line region of over $6\times10^5$\hbox{$\cm^{-3}\K$}\ at 30{\rm\thinspace kpc}\ from the quasar core (Crawford \& Fabian (1989). The quasar again shows a depolarization asymmetry (Garrington et al 1991), and narrow-band imaging by Hes (1995) suggests that the [OII] line emission is extended along the same position angle of $\sim150$$^\circ$ as the strong radio jet to the south-east of the quasar core. The fits to the wobble-corrected profiles show an improvement to the fit with the addition of an extended component, preferably a King model. The lengthscale $R$ of this component lies in the range 5-10 arcsec (where 1 arcsec corresponds to 7.4{\rm\thinspace kpc}\ at the redshift of the quasar), and contains 8-22 per cent of the total X-ray luminosity at 3-8$\times10^{44}$\hbox{$\erg\s^{-1}\,$}. \subsection{H1821+643 } H1821+643 is radio-quiet quasar which is luminous in the infrared. The X-ray emission was found to have a significant extended component by Hall et al (1997). It is included here as a comparison object. The fits improve dramatically with the inclusion of an extended component. With the 50 arcsec apertures, the King model is often preferred, whereas the 100 arcsec apertures show a marked improvement for the broken power-law model. The characteristic radius of this extended component, $R$, shows a wide range of 29-81 arcsec (where 1 arcsec corresponds to 5.5{\rm\thinspace kpc}\ at the redshift of the quasar), but clear separation according to both extended model and outer profile radius employed. The extended component contains from 10--19 per cent of the X-ray luminosity, at 8--20$\times10^{44}$\hbox{$\erg\s^{-1}\,$}. \section{Summary of results} We tabulate average properties of the extended emission component from the profile fits to each quasar in Table~\ref{tab:averages}, where these averages are derived only from the fits where the normalizations of the PSF and model for the extended emission are both allowed to vary (labelled as PSF+broken power-law and PSF+King law in Table~2). The values are averaged from the fits only to the wobble-corrected images (and from each phase interval employed). As is clear from both Tables~2 and \ref{tab:averages}, where the scatter in properties is sufficiently low for comparison to be made (3C48, 3C215, 3C273 and 3C281), systematic differences can be seen between parameters derived using the King and broken power-law fits. The King law always gives a higher X-ray luminosity (and thus higher overall percentage of the total luminosity), but a smaller characteristic radius $R$ than the broken power-law model. The average bolometric luminosities (from the profile fit out to 100 arcsec) span the range of 6-43$\times10^{44}$\hbox{$\erg\s^{-1}\,$}, with values of 13 and 18$\times10^{44}$\hbox{$\erg\s^{-1}\,$} typical for the broken power-law and King models respectively. The King models give values of $R$ in the range 33-38{\rm\thinspace kpc}, with the exception of 14{\rm\thinspace kpc}\ for the fits to 3C273; for the broken power-law model the values are 31-49{\rm\thinspace kpc}. The three quasars, 3C334, 3C254 and 3C275.1, for which we do not differentiate between parameters derived from each model have an average $R$ of 49, 94 and 270 {\rm\thinspace kpc}, respectively. Where a preference between the two models for the extended emission can be seen, it is nearly always for the King law over the broken power-law model (in 3C48, 3C273, 3C281 and 3C334). An exception is the 100 arcsec profile fits to 1821+643, which prefers the broken power-law fits. The continuing increase in luminosity of the extended X-ray emission between the 50 and 100 arcsec profile fits also shows that this component is truly extended over cluster-wide scales of $\sim400-800${\rm\thinspace kpc}\ radius. We note that the discrepancy between the observed, corrected profile and the PSF is least for 3C273 and 3C48, which are the nearest 3CR quasars in our sample. If there are systematic errors associated with the wobble correction that we are unaware of, then these objects will be the most seriously affected. It is important that the extended emission from these objects be confirmed with {\em Chandra}. \addtocounter{table}{+1} \onecolumn \begin{table} \caption{Average properties of the extended component \label{tab:averages}} \begin{tabular}{lcccccc} & & & & & & \\ Quasar & R (kpc) & \% & $L_{bol}$ ($10^{43}$\hbox{$\erg\s^{-1}\,$}) & R (kpc) & \% & $L_{bol}$ ($10^{43}$\hbox{$\erg\s^{-1}\,$}) \\ outer radius (arcsec) & (50) & (50) & (50) & (100) & (100) & (100)\\ & & & & & & \\ 3C48 (King) & 33$\pm$4 & 13$\pm$1 & 154$\pm$13 & 35$\pm$2 & 15$\pm$2 & 180$\pm$20 \\ \ \ $''$ \ \ \ (BPL) & 39$\pm$1 & 11$\pm$1 & 119$\pm$7 & 39$\pm$1 & 11$\pm$1 & 128$\pm$9 \\ 3C215 (King) & 32$\pm$1 & 26$\pm$3 & 85$\pm$11 & 33$\pm$1 & 26$\pm$4 & 89$\pm$13 \\ \ \ $''$ \ \ \ \ \ (BPL) & 49$\pm$3 & 17$\pm$3 & 57$\pm$9 & 49$\pm$3 & 19$\pm$3 & 68$\pm$11 \\ 3C254 & 86$\pm$10 & 16$\pm$2 & 121$\pm$9 & 94$\pm$11 & 17$\pm$1 & 141$\pm$4 \\ 3C273 (King) & 15$\pm$1 & 10$\pm$1 & 353$\pm$22 & 14$\pm$1 & 12$\pm$1 & 426$\pm$17 \\ \ \ $''$ \ \ \ \ \ (BPL) & 31$\pm$2 & 5$\pm$1 & 173$\pm$2 & 31$\pm$2 & 5$\pm$1 & 166$\pm2$\\ 3C275.1 & (341) & (17) & (40) & 270$\pm$7 & 24$\pm$1 & 62$\pm$4 \\ 3C281 (King) & 39$\pm$1 & 62$\pm$6 & 256$\pm$3 & 38$\pm$1 & 64$\pm$4 & 276$\pm$3 \\ \ \ $''$ \ \ \ \ \ (BPL) & 50$\pm$2 & 45$\pm$3 & 189$\pm$8 & 48$\pm$5 & 47$\pm$4 & 213$\pm$2 \\ 3C334 & 50$\pm$10 & 16$\pm$3 & 106$\pm$18 & 49$\pm$9 & 17$\pm$3 & 117$\pm$18 \\ H1821+643 (King) & 175$\pm$12 & 11$\pm$1 & 218$\pm$11 & 267$\pm$18 & 18$\pm$1 & 380$\pm$9 \\ \ \ $''$ \ \ \ \ \ \ \ \ \ \ \ (BPL) & 196$\pm$6 & 13$\pm$1 & 254$\pm$11 & 389$\pm$56 & 20$\pm$1 & 431$\pm$11 \\ \end{tabular} \\ Notes:\\ These values are averaged only from the results to the fits to the wobble-corrected images (and each phase interval employed), and from the PSF+broken power-law and PSF+King law models. Where the two models give consistently different answers, we deduce an average for each (eg 3C48); where there is a similar range or too few fits available to reliably discriminate for such differences (eg 3C275.1) we obtain an average value from both models together. \\ The errors given are the variation in this average value and not a significance of the detection. \\ \end{table} \begin{table} \caption{Derived cooling flow parameters \label{tab:cf}} \begin{tabular}{lccccc} & & & & & \\ Quasar & $L_{\rm CF}$ & \hbox{$\dot M$} & $r_{\rm CF}$ & $P$(30{\rm\thinspace kpc}) & $n(R)$ \\ & ($10^{43}$\hbox{$\erg\s^{-1}\,$}) & (\hbox{$\Msun\yr^{-1}\,$}) & ({\rm\thinspace kpc}) & ($10^6${\rm\thinspace yr}) & ($10^{-2}$\hbox{$\cm^{-3}\,$}) \\ 3C48 & 36 & 330 & 155 & 3.5 & 3.4 \\ 3C215 & 18 & 160 & 120 & 2.2 & 1.7 \\ 3C254 & 45 & 410 & 165 & 2.7 & 1.2 \\ 3C273 & 59 & 510 & 175 & 5.1 & 6.2 \\ 3C275.1& 30 & 100 & 110 & 1.4 & 0.2 \\ 3C281 & 58 & 510 & 180 & 4.0 & 3.0 \\ 3C334 & 33 & 290 & 150 & 3.1 & 2.3 \\ \end{tabular} \\ Notes:\\ $L_{CF}$ is the X-ray luminosity of the extended component within the break radius $R$, and is assumed to be due to a cooling flow in the cluster. \\ \hbox{$\dot M$} is the derived mass cooling rate within radius $R$. \\ $t_{\Lambda}(R)$ is the cooling time at radius $R$.\\ $P$ is the gas pressure at a radius of 30{\rm\thinspace kpc}\ from the centre of the cluster.\\ $n(R)$ is the electron density at radius $R$. \\ \end{table} \twocolumn \section{Discussion} We plot the average bolometric luminosity of the extended cluster component (from Table~\ref{tab:averages}) against quasar redshift in Figure~\ref{fig:lxz}. For comparison we plot the bolometric luminosity of the X-ray source associated with the distant radio galaxies 3C277.2, 3C294, 3C324, 3C356, 3C368 (from Crawford \& Fabian 1996b) and 1138-262 (Carilli et al 1998), and the clusters surrounding the two nearby FR~II radio galaxies Cygnus~A (Ueno et al 1994) and 3C295 (Henry \& Henriksen 1986). The observed countrates of the distant radio galaxies have been converted to luminosities assuming the same 4{\rm\thinspace keV}\ thermal bremsstrahlung model used to obtain luminosities for the quasar extended emission. The luminosities we have derived for the environment of our quasars are brighter than the upper limits of 1.6 -- 3.5$\times10^{44}$\hbox{$\erg\s^{-1}\,$} (rest-frame 0.1-2.4{\rm\thinspace keV}) to any cluster emission surrounding three radio-loud quasars in Hall et al (1995, 1997). We note, however, that those upper limits have been obtained from images {\em not} corrected for satellite wobble. They also assume therefore that the quasar light follows the standard HRI PSF derived by David et al (1995) and accounts for all the light in the innermost bin. Our quasar host clusters are consistent with the luminosity of 3.7$\times10^{45}$\hbox{$\erg\s^{-1}\,$} detected by Hall et al (1997) for the environment of the radio-intermediate quasar H1821+643 at a redshift $z$=0.297. The inferred bolometric luminosities of the extended components we have found here are completely reasonable for moderately rich clusters of galaxies at low redshift. They are comparable to the luminosities of the clusters associated with the powerful radio galaxies Cygnus A and 3C295. They are however (Fig.~1) more luminous than the extended X-ray emission detected around more distant 3CR radio galaxies above redshift one. Whether this indicates evolution, a problem for radio galaxy/quasar unification, or is a result of small number statistics must await the compilation of a complete sample, which are study is not. All extended models have a central cooling time considerably shorter than a Hubble time. We therefore explore the properties of the implied cooling flows occurring around these quasars by deriving some approximate parameters from the broken power-law fits to the profiles. We attribute all the X-ray luminosity of the extended component within radius $R$ to thermal bremsstrahlung from gas with electron density $n$ (where $n\propto r^{-1}$) at a temperature of 4{\rm\thinspace keV}. The cooling time of the gas at $R$ (except in the case of 3C275.1) is then between about 1--3 billion yr. We then estimate the cooling flow radius $r_{\rm CF}$ at which the cooling time is $10^{10}$~yr and obtain a rough indication of the mass deposition rate within that radius from the ratio of the mass of gas within $r_{\rm CF}$ to $10^{10}$~yr. The derived values are shown in Table~\ref{tab:cf}. Note they are of course subject to not only the appropriateness of the fixed slopes chosen for our original broken power-law model, but also to the true gravitational potential of any cluster, and the amount of gravitational work done on the cooling gas. The values should be regarded as uncertain by a least a factor of 2. They may be underestimated by a factor of at least 2 if the gas temperatures are significantly higher than the 4~keV assumed and there is internal absorption such as is common in low redshift cooling flows. Note that the radius of the surface brightness break which we infer is in the range 40--90~kpc, and is similar to the break radius in the profile of the cluster around IRAS~09104 ($\sim60$~kpc, Crawford \& Fabian 1995b). This is likely to be the radius of the core of the gravitational potential of the cluster; the $r^{-1}$ profile then occurs within there since that gas is cooling at approximately constant pressure. Such small gravitational core radii are characteristic of relaxed lensing cluster cores such as are associated with massive cooling flows (Allen 1998). The large break radii found for 3C275.1 and H1821+643 do not agree with this picture and require more detailed images. We can also use our cooling flow parameters to derive a gas pressure $P$ at a radius of 30{\rm\thinspace kpc}\ (see Table \ref{tab:cf}) for comparison to the pressures derived from the completely independent method using the ionization state of the extended optical emission lines (Crawford \& Fabian 1989; Bremer et al 1992; Crawford \& Vanderreist 1997). The pressures derived from the optical nebulosities mostly underestimate those derived from the X-ray profile fits (Figure~\ref{fig:pvsp}) by a factor of up to 10. Given that the gas pressures are mostly derived from the optical nebulosity using a conservative underestimate to the crucial but unknown UV and soft X-ray band of the ionizing nuclear spectrum, this discrepancy is not surprising. Support for this interpretation of the disparity in derived pressures is found in Crawford et al (1991) where a better knowledge of the ionizing continuum of the nucleus of 3C263 was found to increase the optically-derived pressure by up to an order of magnitude. In addition, we note that 3C254, the quasar with the best agreement between the two pressure values is the only one where UV HST data has been used to constrain the shape of the ionizing continuum (Crawford \& Vanderriest 1997). We note that the derived cluster luminosity for 3C273 is high and implies the presence of a rich cluster which is not seen at other wavelengths. At its relatively low redshift of $z=0.16$, such a cluster should be obvious in the optical band. As mentioned already, it has the profile most susceptible to systematic errors, and the absence of an optical cluster may argue for the existence of such errors. If the PSF is then uncertain by a relative amount equal to the observed 3C273 profile and our empirical PSF, then it will not change greatly our results on the other quasars except perhaps for 3C48. Inverse-Compton scattering of quasar radiation could still contribute to our extended component of X-ray emission. If a significant process, the X-ray source would appear asymmetric and lop-sided, as (back) scattering by the electrons in the more distant radio lobe should be stronger than in the nearer lobe. The quality of our current data is insufficient to show any significant asymmetry, and so we are unable to assess the contribution from this process. Observations of these sources with {\em Chandra} and {\em XMM} should clarify its relative importance to the total emission. \begin{figure} \psfig{figure=lxz.ps,width=0.45\textwidth,angle=270} \caption{ \label{fig:lxz} The average bolometric luminosities (as given in Table~\ref{tab:averages}) for the extended component of emission (solid circle markers; values extracted from the 100 arcsec profile fits) plotted against the redshift of the quasar. Separate averages obtained for the same quasar assuming the King or broken power-law models are plotted as two points at the same redshift joined by a straight line (where the King law gives the upper end of the range). The luminosities of the host clusters of the nearby FR~II radio galaxies Cygnus~A and 3C295 are plotted as open circles. The luminosities of the X-ray source associated with the distant radio galaxies 3C277.2, 3C294, 3C324, 3C356, 3C368 and 1138-262 are also plotted (triangle markers), where the observed countrates have been converted to X-ray luminosities assuming the same 4{\rm\thinspace keV}\ thermal bremsstrahlung model used for the quasar extended emission. } \end{figure} \begin{figure} \psfig{figure=pvsp.ps,width=0.45\textwidth,angle=270} \caption{ \label{fig:pvsp} Comparison of the gas pressure at a projected distance of 30{\rm\thinspace kpc}\ from the quasar derived from the ionization state of the extended emission-line gas, and from the broken power-law fits to the X-ray radial profiles. Pressures are expressed in units of $10^5$\hbox{$\cm^{-3}\K$}, and the solid line indicates the locus of equal pressures. } \end{figure} \section{Conclusions} The seven powerful radio-loud quasars studied here appear to be surrounded by luminous extended X-ray emission. The spatial properties of the emission are consistent with an origin in thermal emission from an intracluster medium. The radiative cooling time of the gas within $\sim 50$~kpc of the quasars is only a few billion years or less, indicating the presence of strong cooling flows of hundreds of $\hbox{$\Msun\yr^{-1}\,$}$. The high pressure of that gas is sufficient to support the extended optical nebulosities seen around many of these quasars and may play a r\^ole in shaping the properties of the radio sources, such as structure, depolarization and possibly even fuelling and evolution (Fabian \& Crawford 1990). Such a r\^ole is not clear from a comparison of radio properties with the neighbouring optical galaxy density (Rector et al 1995). A small number of luminous clusters without central AGN have so far been found beyond redshift 0.5, and the discovery of a cluster at $z\sim0.8$ with luminosity $10^{45}$\hbox{$\erg\s^{-1}\,$} (Donahue et al 1998) provides strong evidence for a low density universe. We have shown that the study of the environment of powerful radio-loud quasars is a promising way of extending the discovery of similarly luminous clusters both in numbers and to higher redshifts. \section{Acknowledgements} We thank Steve Allen for advice on the profiles of massive cooling flows. CSC and ACF thank the Royal Society for financial support. This work has been supported in part by the DLR (format DARA GmbH) under grant 50~OR~9403~5 (GH and IL). This research has made use of the NASA/IPAC Extragalactic Database (NED) and the Leicester Database and Archive Service (LEDAS).
\section{Introduction} The investigation of heavy quarks production in high energy hadron collisions provides a method for studying the internal structure of hadrons. Some problems are the same in hadroproduction and photo-/electroproduction processes. So, the review of the situation of the heavy quark hadroproduction can be useful for the interpretation of HERA data. In this talk we present a short review of heavy quark hadroproduction. The theoretical predictions are usually obtained in the NLO parton model \cite{1}. The assumptions which are used for simplifications of the computations are considered in Sect. 2. In the case of one-particle distributions even LO ($\sim \alpha_s^2$) parton model with collinear approximation is enough for the data description, NLO contributions ($\sim \alpha_s^3$) only change the normalizations. On the other hand, in the case of two-particle distributions, see Sect. 3, the collinear approximation has failed, and it is necessary to account for the transverse momenta of the incident partons. The possibility to include the transverse momenta of the incident partons in the framework of semihard theory \cite{GLR}, where the virtualities and polarizations of the gluons are taken into account, is considered in Sect. 4. In Ref. \cite{3} we presented the results for main and simplest subprocess, $gg \rightarrow \overline{Q} Q \; (\sim \alpha_{s}^{2})$ for hadroproduction, and $\gamma g \rightarrow \overline{Q} Q \; (\sim \alpha_{s})$ for photo- and electroproduction. \section{Conventional NLO parton model} The conventional NLO parton model expression for the heavy quark hadroproduction cross sections has the factorization form \cite{CSS}: \begin{equation} \sigma (a b \rightarrow Q\overline{Q}) = \sum_{ij} \int dx_i dx_j G_{a/i}(x_i,\mu_F) G_{b/j}(x_j,\mu_F) \hat{\sigma} (i j \rightarrow Q \overline{Q}) \;, \label{pm} \end{equation} where $G_{a/i}(x_i,\mu_F)$ and $G_{b/j}(x_j,\mu_F)$ are the structure functions of partons $i$ and $j$ in the colliding hadrons $a$ and $b$, $\mu_F$ is the factorization scale (i.e. virtualities of incident partons) and $\hat{\sigma} (i j \rightarrow Q \overline{Q})$ is the cross section of the subprocess which is calculated in perturbative QCD. The last cross section can be written as a sum of LO and NLO contributions, $\hat{\sigma} (i j \rightarrow Q\overline{Q}) = \alpha_s^2(\mu_R) \sigma^{(o)}_{ij} + \alpha_s^3(\mu_R) \sigma^{(1)}_{ij}$, where $\mu_R$ is the renormalization scale, and $\sigma^{(o)}_{ij}$ as well as $\sigma^{(1)}_{ij}$ depend practically only on one variable $\rho = \frac{4m_Q^2}{\hat{s}}$ , $\hat{s} = x_i x_j s_{ab}$. The expression (1) corresponds to the process shown schematically in Fig. 1 with \begin{equation} q_{1T} = q_{2T} =0 \;. \end{equation} The main contribution to the cross section at small $x$ is known to come from gluon-gluon fusion, $i = j = g$. \begin{figure}[htb] \begin{center} \mbox{\psfig{file=tal1a.eps,width=0.35\textwidth}} \\ Fig. 1. Heavy quark production in parton model. \end{center} \end{figure} The principal uncertainties of any numerical QCD calculation of heavy flavour production are connected with the unknown values of the parameters: both scales, $\mu_F$ and $\mu_R$\footnote{These uncertainties should disappear when one sums up all the high order contributions. Sometimes people say that strong scale dependence of the calculated results in LO or NLO means the large contribution of high order diagrams and weak dependence means their small contribution. Of course, it is not true. Strong scale dependence of NLO results means only strong scale dependence of high order contributions but at some fixed scale value the last ones can be numerically small. Weak scale dependence of NLO results means weak scale dependence of high order terms but they can be numerically large.}, and the exact value of heavy quark mass, $m_Q$. The values of both scales should be of the order of hardness of the considered process, however nobody can say what is better to use for scales, $m_Q$, $m_T = \sqrt{m_Q^2 + p_T^2}$ or $\hat{s}$. The phenomenological parton densities are sometimes (at very small $x$) in contradiction \cite{ASS} with the general properties of perturbative QCD. However it is just the region that dominates in the heavy quark production at high energies\footnote{In the case of charm production, $m_c$ = 1.4GeV, at LHC, $\sqrt{s}$ = 14 TeV, the product $x_1x_2$ of two gluons (both $x_1$ and $x_2$ are the integral variable) is equal to $4\cdot 10^{-8}$.}. Another problem of parton model is the collinear approximation. The transverse momenta of the incident partons, $q_{iT}$ and $q_{jT}$ are assumed to be zero, and their virtualities are accounted for only via structure functions; the cross sections $\sigma^{(o)}_{ij}$ and $\sigma^{(1)}_{ij}$ are assumed to be independent on these virtualities. The NLO parton model calculations of the total cross sections of $c\bar{c}$ and $b\bar{b}$ production, as functions of the beam energy, for $\pi^- N$ and $p-N$ collisions can be found in \cite{FMNR}. These results depend strongly (on the level of several times) on the numerical values of quark masses as well as on the both scales, $\mu_F$ and $\mu_R$. Some experimental data are in contradiction with each other, however generally they are in agreement with NLO parton model predictions. \begin{figure}[htb] \begin{center} \mbox{\psfig{file=tal3.eps,width=0.50\textwidth}} \\ Fig. 2. $p_T$-distributions for $p\bar{p} \rightarrow b + X$ at $\sqrt{s}$ = 1.8 TeV, for different values of rapidity. \end{center} \end{figure} The NLO contributions to one-particle distributions lead only to renormalization of LO results, practically without correction of the shapes of a distributions \cite{NDE1,MNR}. It means that instead of more complicate calculation of $p_T$, or rapidity distributions, in NLO, it is enough to calculate them in LO, and multiply after by K-factor \begin{equation} K = \frac{LO + NLO}{LO} \;, \end{equation} which can be taken, say, from the results for total production cross sections. The comparison of LO + NLO calculations with LO multiplied by K-factor is presented in Fig. 2 taken from Ref. \cite{NDE1}. The values of K-factors and their energy and scale dependences for several sets of structure functions were calculated in Refs. \cite{SCG,LSCSh}. The experimental data for $x_F$-distributions of D-mesons produced in $\pi N$ interactions \cite{Adam,Alv} are in agreement with the parton model distributions for bare quarks, as one can see in Fig. 3 taken from \cite{FMNR}. It means that the fragmentation processes are not important here, or they are compensated by, say, recombination processes. The shape of $x_F$-distributions does not depend practically on the mass of $c$-quark. \begin{figure} \begin{center} \centerline{\epsfig{figure=f5a.eps,width=0.45\textwidth,clip=} \hspace{0.3cm} \epsfig{figure=f5b.eps,width=0.45\textwidth,clip=}} Fig. 3. Experimental $x_F$ distributions for $D$ mesons, compared to the NLO parton model prediction for charm quarks. \end{center} \end{figure} The data on one-particle $p_T$-distributions,including the hadronic colliders data for the case of beauty production, also can be described by the NLO parton model, see \cite{FMNR}. \section{Azimuthal correlations and failure of the collinear approximation} The azimuthal angle $\phi$ is defined as an opening angle between two produced heavy quarks, projected onto a plane perpendicular to the beam. In the LO parton model this angle between them is exactly $180^o$. In the case of NLO parton model a distribution over $\phi$ angle appears \cite{MNR}. The investigation of such distributions is very important. In one-partricle distributions, the sum of LO and NLO contributions of the parton model practically coinsides with the LO contribution multiplied by $K$-factor. So we can not control the magnitudes of LO and NLO contributions separately. In the case of azimuthal correlations all difference from the trivial $\delta(\phi - \pi)$ distribution comes from NLO contribution. The experimental data on azimuthal correlations are claimed (see \cite{BEAT} and Refs. therein) to be in disagreement with the NLO predictions, for the cases of charm pair hadro- and photoproduction at fixed target energies. The level of disagreements can be seen in Fig. 4 (solid histograms) taken from Ref. \cite{FMNR}. These data can be described \cite{FMNR}, assuming the comparatively large intrinsic transverse momenta of incoming partons ($k_T$ kick). For each event, in the longitudinal centre-of-mass frame of the heavy quark pair, the $Q\overline{Q}$ system is boosted to rest. Then a second transverse boost is performed, which gives the pair a transverse momentum equal to $\vec{p}_T(Q\overline{Q}) + \vec{k}_T(1) + \vec{k}_T(2)$; $\vec{k}_T(1)$ and $\vec{k}_T(2)$ are the transverse momenta of the incoming partons, which are chosen randomly, with their moduli distributed according to \begin{equation} \frac{1}{N}\frac{dN}{dk_T^2}=\frac{1}{\langle k_T^2 \rangle} \exp(-k_T^2/\langle k_T^2 \rangle). \end{equation} \begin{figure} \begin{center} \centerline{\epsfig{figure=f7a.eps,width=0.45\textwidth,clip=} \hspace{0.3cm} \epsfig{figure=f7b.eps,width=0.45\textwidth,clip=}} Fig. 4. Azimuthal correlation for charm production in $\pi N$ collisions: NLO parton model and $k_T$ kick calculations versus the WA75 and WA92 data. \end{center} \end{figure} The dashed and dotted histograms in Fig. 4 correspond to the NLO parton model prediction, supplemented with the $k_T$ kick, with $\langle k_T^2\rangle=0.5$ GeV$^2$ and $\langle k_T^2\rangle=1$ GeV$^2$, respectively. We see that with $\langle k_T^2 \rangle= 1$ GeV$^2$ it is possible \cite{FMNR} to describe the data. However the large intrinsic transverse momentum significantly changes one-particle $p_T$-distri\-butions of heavy flavour hadrons, which were in good agreement with the data. The solid curves in Fig. 5 taken from \cite{FMNR} represent the NLO parton model predictions for charm quarks $p_T$-distributions which are in agreement with the data. The effect of the $k_T$ kick results in a hardening of the $p_T^2$ spectrum. On the other hand, by combining the $k_T$ kick with $\langle k_T^2 \rangle =1$ GeV$^2$ and the Peterson fragmentation \cite{Pet}, the theoretical predictions slightly undershoot the data (dot-dashed curves). \begin{figure} \begin{center} \centerline{\epsfig{figure=f4a.eps,width=0.45\textwidth,clip=} \hspace{0.3cm} \epsfig{figure=f4b.eps,width=0.45\textwidth,clip=}} Fig. 5. Charm $p_T^2$ distribution measured by WA92 and E769, compared to the NLO parton model predictions, with and without the non-perturbative effects. \end{center} \end{figure} The $k_T$ kick can only very weakly change the $x_F$-distributions of produced $c$-quarks, Fig. 3, and after accounting the fragmentation these distributions can become too soft. Let us consider why the conventional NLO parton model with collinear approximation works reasonably for one-particle diustributions, and, at the same time, it is in disagreement with the data on azimuthal correlations. The contribution of the processes of Fig. 1, which governs the heavy quark production can be written\footnote{We omit for simplicity all factors which are non-essential here.} as a convolution of initial transverse momenta distributions, $I(q_{1T})$ and $I(q_{2T})$, with squared modulo of perturbative QCD matrix element, $\vert M(q_{1T}, q_{2T}, p_{1T}, p_{2T})\vert ^2$ : \begin{equation} \sigma_{QCD}(Q\overline{Q}) \propto \int d^2 q_{1T} d^2 q_{2T} I(q_{1T}) I(q_{2T}) \vert M(q_{1T}, q_{2T}, p_{1T}, p_{2T}) \vert ^2 \;. \end{equation} Now there are two possibilities: \\ i) the essential values of initial transverse momenta are much smaller than the transverse momenta of produced heavy quarks, $q_{iT} \ll p_{iT}$, and \\ ii) all transverse momenta are of the same order, $q_{iT} \sim p_{iT}$. In LO for the first case (i) we can wait that one-dimentional $p_T$ distributions should be more broad than the distributions on the transverse momenta of the quark pair, because in LO $q_{1T} + q_{2T} = p_{1T} + p_{2T}$. NLO gives here a correction, numerically not very large, see Fig. 2. In this case one can replace both the initial distributions $I(q_{iT})$ by $\delta$-functions, $\delta(q_{iT})$. This reduces the expression (5) to the very simplified one : \begin{equation} \sigma_{coll.}(Q\overline{Q}) \propto \vert M(0, 0, p_{1T}, p_{2T}) \vert ^2 \;, \end{equation} in total agreement with Weizsaecker-Williams approximation in QED. In the second case (ii) the distributions on the transverse momenta of the quark pair should be the same, or even more broad than the one-particle $p_T$-distributions, and namely this situation is realised \cite{FMNR}. In this case we can not wait a priory that the Weizsaecker-Williams approximation will give good results, however it works quite reasonably in the case of one-particle distributions. In the case of distribution on the transverse momentum of the heavy-quark pair, $p_T^2(Q\overline{Q})$ we measure (only approximately in NLO) the distribution over the sum of transverse momenta of incident gluons. In this case it seems to be senseless to replace the distributions $I(q_{1T})$ and $I(q_{2T})$ by $\delta$-functions, and to expect a reasonable agreement with the data. The same can be said about the azimuthal correlations. The $k_t$ kick \cite{FMNR} effectively accounts for the transverse momenta of incident partons. It uses the expression which can be written symbollically as \begin{equation} \sigma_{kick}(Q\overline{Q}) \propto I(q_{1T}) I(q_{2T}) \otimes \vert M(0, 0, p_{1T}, p_{2T}) \vert ^2 \;, \end{equation} and the main difference from the general QCD expression Eq. (5) is that due to absence of $q_{iT}$ in the matrix element the values of $\langle k_T^2\rangle$ in Eq. (4) should be different for different processes and kinematical regions. The reason is that in Eq. (5) the values $I(q_{iT})$ decrease at large $q^2_{iT}$ as a weak power (see next Sect.), i.e. comparatively slowly, and more important is the $q^2_{iT}$ dependence of the matrix element. In the last one the corrections of the order of $q^2_{iT}/\mu^2$, where $\mu^2$ is the QCD scale, are small enough when $q^2_{iT}/\mu^2 << 1$ and they start to suppress a matrix element value when $q^2_{iT}/\mu^2 \sim 1$. \section{Heavy quark production in semihard approximation} Let us consider another approach, when the transverse momenta of incident gluons in the small-$x$ region appear from the diffusion of transverse momenta in the gluon evolution\footnote{The similar approach based on $k_T$-factorization formulae can be found, for example, in Refs. \cite{CCH,CE,MW}.}. This diffusion is described by the function $\varphi(x,q^2)$ determined \cite{GLR} as \begin{equation} \label{xg} \varphi (x,q^2) = 4\sqrt{2}\,\pi^3 \frac{d\,[xG(x,q^2)]}{d q^2} \;, \end{equation} where $G(x,q^2)$ is usual gluon structure function. In principle the function $\varphi(x,q^2)$ which determine the probability to find gluon with fixed value of transverse momentum, $q_T$, depends on three variables, $x$, $q_T$ and gluon virtuality $q^2$. However at small $x$ in LLA $q_T^2 \approx - q^2$, and it leads to comparatively weak dependence of $\varphi(x,q^2)$ on $q_T^2$ (strongly different from exponential dependence in Eq. (4)) due to weak $q^2$-dependences of phenomenological structure functions. The exact expression for gluon $q_T$-distributions can be obtained, as a solution of the nonlinear evolution equation. The calculations \cite{Blu} result in difference from our $\varphi(x,q^2)$ function only about 10-15\%. The matrix element $M_{QQ}$ accounting for the gluon virtualities and polarizations is much more complicate than the parton model one. That is why we consider only LO contribution of the subprocess $gg \to Q\bar{Q}$. The differential cross section of heavy quarks hadroproduction has the form $$ \frac{d\sigma_{pp}}{dy^*_1 dy^*_2 d^2 p_{1T}d^2 p_{2T}}\,=\,\frac{1}{(2\pi)^8} \frac{1}{(s)^2}\int\,d^2 q_{1T} d^2 q_{2T} \delta (q_{1T} + q_{2T} - p_{1T} - p_{2T}) $$ \begin{equation} \label{spp} \times\,\,\frac{\alpha_s(q^2_1)}{q_1^2} \frac{\alpha_s (q^2_2)}{q^2_2} \varphi(q^2_1,y)\varphi (q^2_2, x)\vert M_{QQ}\vert^2. \end{equation} Here $s = 2p_a p_b\,\,$ and $y^*_{1,2}$ are the quarks' rapidities in the hadron-hadron c.m.s. frame. Eq. (9) enables to calculate straightforwardly all distributions concerning heavy flavour one-particle, or pair production. However there exists a problem coming from infrared region. Gluon structure function in Eq. (8) is not determined at small virtualities, so the function $\varphi (x,q^2_2)$ is unknown at the small values of $q^2_2$ and $q^2_1$. To solve this problem we will use the direct consequence of Eq. (8) \cite{Kwi} \begin{equation} xG(x,q^2) = xG(x,Q_0^2) + \frac{1}{4\sqrt{2}\,\pi^3} \int_{Q_0^2}^{q^2} dq_1^2 \varphi (x,q_1^2) \;, \end{equation} and rewrite \cite{3} the integrals in the Eq. (9) as the sum of four contributions. The first one is determined by the product of two gluon distributions, $G(x,Q_0^2)$ and $G(y,Q_0^2)$, and it is the same as the conventional LO parton model expression. Next three terms contain the corrections to the parton model. If the initial energy is not high enough, the first term dominates. In the case of very high energy the first term can be considered as a small corrections, and our results are differ from the conventional ones. In the cases when the collinear approximation is available, our results only slightly differ from the parton model. \begin{figure}[htb] \begin{center} \mbox{\psfig{file=tal6.eps,width=0.50\textwidth}} \\ Fig. 6. Cross section of beauty production in CDF. \end{center} \end{figure} It is illustrated in Fig. 6 taken from \cite{3}, where we compare our calculations of $b$-quark $p_T$-distributions with the experimental results of CDF collaboration and with two variants of parton model calculations. The distributions over the azimuthal angle $\phi$ can be found in \cite{SS}. The essential values of $q_{1T}$ and $q_{2T}$ in our calculations increase with increase the value of $p_T^{min}$ of detected $b$-quark. In the language of $k_T$ kick it means that the values of $\langle k_T^2\rangle$ will be also increased. \section{Conclusion} The experimental results on total cross sections for charm and beauty production are in agreement with the conventional parton model predictions, using reasonable values of QCD scales and quark masses. The data on $x_F$ and $p_T$ distributions are also in reasonable agreement with parton model without any fragmentation functions\footnote{The $p_T$ distributions of charm photoproduction measured by E691 Coll. \cite{E691a} are more hard than the NLO parton model predictions \cite{FMNR,Man97}, that can be considered \cite{Man97} as an argument for including a non-perturbative fragmentation function. However, the $p_T$ slope of the calculated spectrum depends strongly on the charm quark mass, and the measured charm production cross section has very strange energy dependence \cite{E691a}.}. Moreover, the shapes of one-particle LO and NLO distributions practically coinside. It means that instead of calculation the NLO contributions, it is enough to calculate only LO contributions, and rescale them using K-factor taken, say, from the calculated ratio of total cross sections. In the case of distribution over the total transverse momentum of the produced quark pair, or azimuthal correlations, the conventional NLO parton model with collinear approximation can not describe the data. The $k_T$ kick \cite{FMNR} allows one to describe these data, however the problems with one-particle $p_T$-distributions appear, which can be solved by introducing the fragmentation function. The last way should produce the problems in description of $x_F$-spectra. Moreover, it seems that the $\langle k_T^2\rangle$ values should depend on the process and the kinematical regions. Another possibility to solve the problems of initial transverse momenta is to use semihard theory, accounting for the virtual nature of the interacting gluons, as well as their transverse motion and different polarizations. It results in a qualitative differences with the LO parton model predictions \cite{3,SS}. In Ref. \cite{RSS1} the values of $F_2(x,Q^2)$ were calculated using phenomenological gluon structure functions, and the infrared contributions to $F_2(x,Q^2)$ were investigated in details. The possible estimations of the shadow corrections in the processes of heavy flavour production can be found in Refs. \cite{3,LRS}. I am grateful to M.G.Ryskin and A.G.Shuvaev for multiple discussions, to M.L.Mangano for very useful critical comments and to E.M.Levin who participated at the early stage of this activity. I thank the Organizing Commitee of HERA Monte Carlo Workshop for financial support. This work is supported by grant NATO OUTR.LG 971390.
\section{Introduction} The summation of logarithms of $1/x$ in deep inelastic structure functions at small values of Bjorken $x$ leads to the Balitskii-Fadin-Kuraev-Lipatov (BFKL) equation \cite{BFKL,FR}, which in the leading approximation sums terms of order $[\alpha_{\mbox{\tiny S}}\ln(1/x)]^n$. Recently the next-to-leading terms have also been computed \cite{NLOA,NLOB}. Since the dynamics of the small-$x$ region is supposed to be different from that at higher $x$ in several respects, it is important to make and test predictions of a wide range of observables in this region. In recent papers \cite{FSV,Web98} predictions were presented for the rates of emission of fixed numbers of `resolved' final-state gluons, together with any number of unresolvable ones. Here `resolved' means having a transverse momentum larger than some fixed value $\mu_{\mbox{\tiny R}}$. The predictions were valid in the double-logarithmic (DL) approximation, i.e.\ retaining only terms of the form $[\alpha_{\mbox{\tiny S}}\ln(1/x)\ln(Q^2/\mu_{\mbox{\tiny R}}^2)]^n$. In this approximation, each resolved gluon can be equated to a single jet, since to resolve it into more than one jet would cost extra powers of $\alpha_{\mbox{\tiny S}}$ with no corresponding powers of $\ln(1/x)$. The present paper extends the work of ref.~\cite{Web98} to include terms with fewer powers of $\ln(Q^2/\mu_{\mbox{\tiny R}}^2)$, i.e.\ those of the form $[\alpha_{\mbox{\tiny S}}\ln(1/x)]^n\,[\ln(Q^2/\mu_{\mbox{\tiny R}}^2)]^m$ where $0<m<n$, which we refer to as single-logarithmic (SL) corrections. The fact that we still demand a factor of $\ln(1/x)$ with each power of $\alpha_{\mbox{\tiny S}}$ means that the identification of resolved gluons with jets\footnote{When counting jets in deep inelastic lepton scattering, we always omit final-state hadrons that originate from the quark-antiquark pair which couples the gluon to the virtual photon.} remains valid. There are two alternative methods for the calculation of final-state properties at small $x$: the original multi-Regge BFKL method and the CCFM \cite{C,CFM,M} approach, which takes account of the coherence of soft gluon emission. It has been shown, at the DL level in refs.~\cite{FSV,Web98} and now at the SL level \cite{Sal99}, that the two methods are equivalent for the observables considered here. We therefore adopt the BFKL approach, which is calculationally simpler. The paper is organized as follows. In sect.~\ref{sec_bfkl} we recall the BFKL formalism and the predicted behaviour of the gluon structure function at small $x$. In sect.~\ref{sec_rates} we first compute the single-jet rate to SL accuracy and show that the only modification to the DL result comes from the SL corrections to the anomalous dimension. Next, in subsect.~\ref{sec_2jet}, we calculate the SL corrections to the two-jet rate, first in the form of a numerical integral and then as a perturbation series. In subsect.~\ref{sec_3jet} we apply the same methods to the three-jet rate. In sect.~\ref{sec_njet} we derive the SL perturbative expansion of the generating function for the multi-jet rates, and use this to obtain the corresponding expansions for the mean jet multiplicity and its dispersion. Our conclusions are presented in sect.~\ref{sec_conc}. A useful class of integrals is evaluated to SL accuracy in the appendix. \section{BFKL formalism}\label{sec_bfkl} We start from the unintegrated structure function of a single gluon, $f(x,k^2,\mu^2)$, which in the exclusive form of the BFKL approach satisfies the equation \begin{equation}\label{fxk2} f(x,k^2,\mu^2) = \delta(1-x)\,\delta^2(k) + \bar\as\int_{\mu^2}\frac{dq^2}{q^2}\frac{d\phi}{2\pi} \frac{dz}{z^2}\Delta(z,k^2,\mu^2)\,f(x/z,|q+k|^2,\mu^2)\;. \end{equation} Here $\bar\as=3\alpha_{\mbox{\tiny S}}/\pi$, $k$ is the (2-vector) transverse momentum of the gluon probed in the deep inelastic scattering, $q$ is that of an emitted gluon, $\phi$ is the azimuthal angle giving the direction of $q$, $\mu$ is a collinear cutoff and $\Delta$ is the Regge form factor \begin{equation} \Delta(z,k^2,\mu^2) = \exp\left(-\bar\as\ln\frac{1}{z}\ln\frac{k^2}{\mu^2}\right)\;. \end{equation} To carry out the $z$ integration it is convenient to use a Mellin representation, \begin{equation}\label{mellin} f_\omega(k^2,\mu^2) = \int_0^1 dx\,x^\omega f(x,k^2,\mu^2)\;, \end{equation} with inverse \begin{equation}\label{melinv} f(x,k^2,\mu^2) = \frac{1}{2\pi i}\int_C d\omega\,x^{-\omega-1} f_\omega(k^2,\mu^2)\;, \end{equation} where the contour $C$ is parallel to the imaginary axis and to the right of all singularities of the integrand. This gives \begin{equation}\label{bfkl} f_\omega(k^2,\mu^2) = \delta^2(k) + H_\omega(k^2,\mu^2) \int_{\mu^2}\frac{dq^2}{q^2}\frac{d\phi}{2\pi}f_\omega(|q+k|^2,\mu^2) \end{equation} where \begin{equation}\label{delom} H_\omega(k^2,\mu^2)= \frac{\bar\as}{\omega+\bar\as\ln(k^2/\mu^2)}\;. \end{equation} To solve eq.~(\ref{bfkl}) one can make use of the relation, derived in the appendix, \begin{equation}\label{intf} \int_{\mu^2}^{Q^2}\frac{dq^2}{q^2}\frac{d\phi}{2\pi}f(|q+k|^2) = \int_{\mu^2}^{Q^2}\frac{dq^2}{q^2}f(\max\{q^2,k^2\}) +2\sum_{m=1}^\infty \zeta(2m+1) \left(k^2\frac{\partial}{\partial k^2}\right)^{2m}f(k^2)\;, \end{equation} which is valid for $\mu^2<k^2<Q^2$ to logarithmic accuracy, i.e.\ neglecting terms suppressed by powers of $\mu^2/k^2$ or $k^2/Q^2$. Then for $k^2>\mu^2$ the solution is of the form \begin{equation}\label{fomsol} f_\omega(k^2,\mu^2) = \frac{\gamma}{\pi k^2}\left(\frac{k^2}{\mu^2}\right)^\gamma \end{equation} where \begin{equation} 1=H_\omega(k^2,\mu^2)\left(\ln\frac{k^2}{\mu^2}-\frac{1}{\gamma-1} +2\sum_{m=1}^\infty \zeta(2m+1)(\gamma-1)^{2m}\right) \end{equation} and hence \begin{equation}\label{omlip} \omega = -\bar\as\left[2\gamma_{\mbox{\tiny E}}+\psi(\gamma)+\psi(1-\gamma)\right]\;, \end{equation} $\psi$ being the digamma function and $\gamma_{\mbox{\tiny E}}=-\psi(1)$ the Euler constant. The solution of eq.~(\ref{omlip}) is $\gamma=\gamma_{\mbox{\tiny L}}(\alb/\om)$, the Lipatov anomalous dimension: \begin{equation}\label{glip} \gamma_{\mbox{\tiny L}}(\alb/\om)= \frac{\alb}{\om}+ 2\zeta(3)\left(\frac{\alb}{\om}\right)^4+ 2\zeta(5)\left(\frac{\alb}{\om}\right)^6+ 12[\zeta(3)]^2\left(\frac{\alb}{\om}\right)^7+\ldots\;. \end{equation} The integrated gluon structure function at scale $Q^2$ is then given by \begin{equation}\label{FomQ} F_\omega(Q^2,\mu^2) = 1+\pi\int_{\mu^2}^{Q^2}dk^2\,f_\omega(k^2,\mu^2) =\left(\frac{Q^2}{\mu^2}\right)^{\gamma_{\mbox{\tiny L}}(\alb/\om)}\;. \end{equation} Since we are interested in final states, we shall need to decompose the structure function into terms corresponding to different numbers of emitted gluons. The contribution to $f(x,k^2,\mu^2)$ from emission of $n$ gluons is obtained by iteration of eq.~(\ref{fxk2}): \begin{equation} f^{(n)}(x,k^2,\mu^2) = \prod_{i=1}^n\int_{\mu^2} \frac{dq_i^2}{q_i^2}\frac{d\phi_i}{2\pi}\frac{dz_i}{z_i} \bar\as\Delta(z_i,k_i^2,\mu^2)\delta(x-x_n)\delta^2(k-k_n)\;, \end{equation} where \begin{equation} x_i=\prod_{l=1}^i z_l\;,\>\>\>\>k_i=-\sum_{l=1}^i q_l\;. \end{equation} The contribution to the structure function at scale $Q$ is then obtained by integrating over all $\mu^2<q_i^2<Q^2$: \begin{equation}\label{Fn} F^{(n)}(x,Q^2,\mu^2) = \prod_{i=1}^n\int_{\mu^2}^{Q^2} \frac{dq_i^2}{q_i^2}\frac{d\phi_i}{2\pi} \frac{dz_i}{z_i}\bar\as\Delta(z_i,k_i^2,\mu^2)\delta(x-x_n)\;, \end{equation} or in terms of the Mellin transform \begin{equation}\label{Fomn} F^{(n)}_\omega(Q^2,\mu^2) = \prod_{i=1}^n\int_{\mu^2}^{Q^2} \frac{dq_i^2}{q_i^2}\frac{d\phi_i}{2\pi}H_\omega(k_i^2,\mu^2)\;. \end{equation} \section{Jet rates}\label{sec_rates} \subsection{Single-jet rate}\label{sec_1jet} Consider first the effect of requiring one emitted gluon, say the $j$th, to have $q_j^2>\mu_{\mbox{\tiny R}}^2$ while all the others have $q_i^2<\mu_{\mbox{\tiny R}}^2$. This defines the contribution of one resolved gluon plus $n-1$ unresolved, $F^{(n,1\,{\mbox{\scriptsize jet}})}$: \begin{equation} F_\omega^{(n,1\,{\mbox{\scriptsize jet}})}(Q^2,\mu_{\mbox{\tiny R}}^2,\mu^2)= \sum_{j=1}^n \int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_j^2}{q_j^2}\frac{d\phi_j}{2\pi} H_\omega(k_j^2,\mu^2)\prod_{i\neq j}^n\int_{\mu^2}^{\mu_{\mbox{\tiny R}}^2} \frac{dq_i^2}{q_i^2}\frac{d\phi_i}{2\pi}H_\omega(k_i^2,\mu^2)\;. \end{equation} Notice that for $i<j$ the contribution is identical to the $(j-1)$-gluon contribution to the structure function evaluated at $Q^2=\mu_{\mbox{\tiny R}}^2$. On the other hand for $i>j$ we have $k_i^2\simeq q_j^2>\mu_{\mbox{\tiny R}}^2$. As shown in the appendix, when $q_j^2>\mu_{\mbox{\tiny R}}^2$ we can write for any function $f$, to logarithmic accuracy, \begin{equation}\label{Hint1} \int_{\mu^2}^{\mu_{\mbox{\tiny R}}^2}\frac{dq_i^2}{q_i^2}\frac{d\phi_i}{2\pi} f(|q_i+q_j|^2) = \ln\left(\frac{\mu_{\mbox{\tiny R}}^2}{\mu^2}\right)\,f(q_j^2)\;. \end{equation} Thus the $q_i$ integrations for $i>j$ become trivial and \begin{equation}\label{Fn1resom} F^{(n,1\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2,\mu^2) = \frac{1}{S} \sum_{j=1}^n F^{(j-1)}_\omega(\mu_{\mbox{\tiny R}}^2,\mu^2) \int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_j^2}{q_j^2} \left[S\,H_\omega(q_j^2,\mu^2)\right]^{n-j+1} \end{equation} where we define\footnote{Note that we define $S$ and $T$ differently from Refs.~\cite{FSV,Web98} (twice as large) in order to simplify expressions for the single-logarithmic terms.} \begin{equation}\label{STdef} S=\ln(\mu_{\mbox{\tiny R}}^2/\mu^2)\;,\>\>\>\>T=\ln(Q^2/\mu_{\mbox{\tiny R}}^2)\;. \end{equation} Summing over all $j$ and $n$ gives the total one-jet contribution, \begin{eqnarray}\label{F1res} F^{(1\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2,\mu^2) &=& F_\omega(\mu_{\mbox{\tiny R}}^2,\mu^2)\, \int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_j^2}{q_j^2} H_\omega(q_j^2,\mu_{\mbox{\tiny R}}^2)\nonumber\\ &=&\exp[\gamma_{\mbox{\tiny L}}(\alb/\om) S]\,G_\omega^{(1)}(T) \end{eqnarray} where \begin{equation}\label{G1def} G_\omega^{(1)}(T)=\ln\left(1+\frac{\alb}{\om} T\right)\;. \end{equation} Notice that the collinear-divergent part (the $S$-dependence) factorizes, and the fraction of events with one jet is given by the cutoff-independent function \begin{equation}\label{R1res} R^{(1\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2) = \frac{F^{(1\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2,\mu^2)}{F_\omega(Q^2,\mu^2)} \>=\>\exp[-\gamma_{\mbox{\tiny L}}(\alb/\om) T]\,G_\omega^{(1)}(T)\;. \end{equation} Thus in the case of the single-jet rate, the only subleading logarithms are those generated by the presence of the full Lipatov anomalous dimension (\ref{glip}) in eq.~(\ref{R1res}). To obtain the jet cross section as a function of $x$, we note that eq.~(\ref{F1res}) implies that \begin{equation}\label{F1FG} F^{(1\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2) = F_\omega(\mu_{\mbox{\tiny R}}^2)\, G_\omega^{(1)}(T)\;, \end{equation} where we have used the factorization property to replace the cutoff-dependent gluon structure function $F_\omega(\mu_{\mbox{\tiny R}}^2,\mu^2)$ by the measured structure function of the target hadron at scale $\mu_{\mbox{\tiny R}}^2$, $F_\omega(\mu_{\mbox{\tiny R}}^2)$. It follows that the single-jet contribution as a function of $x$ is given by the convolution \begin{equation}\label{F1x} F^{(1\,{\mbox{\scriptsize jet}})}(x,Q^2,\mu_{\mbox{\tiny R}}^2) = F(x,\mu_{\mbox{\tiny R}}^2)\otimes G^{(1)}(x,T) \equiv\int_x^1\frac{dz}{z} F(z,\mu_{\mbox{\tiny R}}^2)\otimes G^{(1)}(x/z,T) \end{equation} where the inverse Mellin transformation (\ref{melinv}) applied to eq.~(\ref{G1def}) gives \begin{equation}\label{G1x} G^{(1)}(x,T)= \frac{1}{2\pi i}\int_C d\omega\,x^{-\omega-1}\,G_\omega^{(1)}(T) \>=\>\frac{1-x^{\bar\as T}}{x\ln(1/x)}\;. \end{equation} \subsection{Two-jet rate}\label{sec_2jet} Now suppose we resolve two gluons $j,j'$ ($j<j'$) with transverse momenta $q_j^2,q_{j'}^2>\mu_{\mbox{\tiny R}}^2$. In place of eq.~(\ref{F1res}) we have \begin{equation}\label{F2res} F^{(2\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2,\mu^2) = F_\omega(\mu_{\mbox{\tiny R}}^2,\mu^2)\, \int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_j^2}{q_j^2}H_\omega(q_j^2,\mu_{\mbox{\tiny R}}^2) \int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_{j'}^2}{q_{j'}^2} \frac{d\phi_{j'}}{2\pi}K_\omega(|q_j+q_{j'}|^2,\mu_{\mbox{\tiny R}}^2) \end{equation} where, defining the dijet transverse momentum $q_J=q_j+q_{j'}$, \begin{equation} K_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)= H_\omega(q_J^2,\mu^2)\left(1+\sum_{n=j'+1}^\infty \prod_{i=j'+1}^n\int_{\mu^2}^{\mu_{\mbox{\tiny R}}^2}\frac{dq_i^2}{q_i^2} \frac{d\phi_i}{2\pi}H_\omega(k_i^2,\mu^2)\right)\;. \end{equation} When $q_J^2>\mu_{\mbox{\tiny R}}^2$ we can safely set $k_i^2=q_J^2$ for $i>j'$, to obtain \begin{equation} K_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)=H_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)\;. \end{equation} However, this cannot be correct for $q_J^2<\mu_{\mbox{\tiny R}}^2$, because $K_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)$ would then be infinite at $q_J^2=\mu_{\mbox{\tiny R}}^2\exp(-\omega/\bar\as)$. Since $K_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)$ must be independent of the cutoff $\mu$, we can evaluate it for $q_J^2<\mu_{\mbox{\tiny R}}^2$ by setting $\mu^2=q_J^2$, which gives \begin{equation}\label{HqltR} K_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)=\frac{\alb}{\om} F_\omega(\mu_{\mbox{\tiny R}}^2,q_J^2) =\frac{\alb}{\om}\left(\frac{\mu_{\mbox{\tiny R}}^2}{q_J^2}\right)^{\gamma_{\mbox{\tiny L}}(\alb/\om)}\;. \end{equation} Thus we can define the continuous function \begin{equation}\label{Kom} K_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)=H_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)\theta(q_J^2-\mu_{\mbox{\tiny R}}^2) +\frac{\alb}{\om} F_\omega(\mu_{\mbox{\tiny R}}^2,q_J^2)\theta(\mu_{\mbox{\tiny R}}^2-q_J^2)\;. \end{equation} The two-jet rate is then given by \begin{equation}\label{R2res} R^{(2\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2) = \frac{F^{(2\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2,\mu^2)}{F_\omega(Q^2,\mu^2)} \>=\>\exp[-\gamma_{\mbox{\tiny L}}(\alb/\om) T]\,G_\omega^{(2)}(T) \end{equation} where \begin{equation}\label{G2def} G_\omega^{(2)}(T) =\int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_j^2}{q_j^2}\frac{dq_{j'}^2}{q_{j'}^2} \frac{d\phi_{j'}}{2\pi}K_\omega(q_j^2,\mu_{\mbox{\tiny R}}^2)\,K_\omega(|q_j+q_{j'}|^2,\mu_{\mbox{\tiny R}}^2)\;. \end{equation} The integrations in eq.~(\ref{G2def}) can be performed numerically, without encountering any non-integrable divergences or discontinuities in the integrand. The resulting two-jet rate is shown by the solid curve in fig.~\ref{fig_R2a} for a relatively small value of $\bar\as/\omega$ (0.2), and for a larger value (0.4) in fig.~\ref{fig_R2b}. The dashed curves show the result of using the DL prediction for $G_\omega^{(2)}$ in eq.~(\ref{R2res}), i.e. \begin{equation}\label{R2DL} R^{(2\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2) \simeq \exp[-\gamma_{\mbox{\tiny L}}(\alb/\om) T]\,G_\omega^{(2,{\mbox{\scriptsize DL}})}(T) \end{equation} where \cite{Web98} \begin{equation}\label{G2DL} G_\omega^{(2,{\mbox{\scriptsize DL}})}(T) = \frac{1}{2}\ln^2\left(1+\frac{\alb}{\om} T\right) +\ln\left(1+\frac{\alb}{\om} T\right) -\frac{\alb}{\om} T\left(1+\frac{\alb}{\om} T\right)^{-1}\;. \end{equation} We see that for $\bar\as/\omega=0.2$ the single-logarithmic correction is small, while for the larger value it is substantial. \FIGURE{\input{full2.pstex_t} \caption{\label{fig_R2a} Two-jet rate for $\bar\as/\omega = 0.2$. Dashed: double-log approximation. Solid: with single-log corrections.}} \FIGURE{\input{full4.pstex_t} \caption{\label{fig_R2b} Two-jet rate for $\bar\as/\omega = 0.4$. Dashed: double-log approximation. Solid: with single-log corrections.}} Next we consider the perturbative expansion of the two-jet rate. We can use eq.~(\ref{Kom}) and the results in the appendix to obtain \begin{eqnarray}\label{Kint} \int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_{j'}^2}{q_{j'}^2}\frac{d\phi_{j'}}{2\pi} K_\omega(|q_j+q_{j'}|^2,\mu_{\mbox{\tiny R}}^2)&=& \int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_{j'}^2}{q_{j'}^2} H_\omega(\max\{q_j^2,q_{j'}^2\},\mu_{\mbox{\tiny R}}^2)\nonumber\\ &+&2\sum_{m=1}^\infty(2m)!\,\zeta(2m+1) \,[H_\omega(q_j^2,\mu_{\mbox{\tiny R}}^2)]^{2m+1}. \end{eqnarray} Substituting in eq.~(\ref{G2def}) we find \begin{equation}\label{G2res} G_\omega^{(2)}(T) = G_\omega^{(2,{\mbox{\scriptsize DL}})}(T) + G_\omega^{(2,{\mbox{\scriptsize SL}})}(T) \end{equation} where $G_\omega^{(2,{\mbox{\scriptsize DL}})}$ is the double-logarithmic result (\ref{G2DL}) and the single-logarithmic correction is \begin{equation}\label{G2SL} G_\omega^{(2,{\mbox{\scriptsize SL}})}(T) = 2\sum_{m=1}^\infty\frac{(2m)!}{2m+1}\zeta(2m+1) \left(\frac{\alb}{\om}\right)^{2m+1} \left[1-\left(1+\frac{\alb}{\om} T\right)^{-2m-1}\right]\;. \end{equation} Note that the series in eq.~(\ref{G2SL}) is strongly divergent for any value of $\bar\as/\omega$. This is due to the singularity of $H_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)$ at $q_J^2=\mu_{\mbox{\tiny R}}^2\exp(-\omega/\bar\as)<\mu_{\mbox{\tiny R}}^2$. The change in $K_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2)$ when $q_J^2<\mu_{\mbox{\tiny R}}^2$ (see eq.~(\ref{Kom})) removes the singularity from the integrand in eq.~(\ref{G2def}) but this does not affect the region of convergence of the perturbation series. The situation is analogous to the way in which the running of the QCD coupling $\alpha_{\mbox{\tiny S}}(q^2)$ at low values of $q^2$ produces infrared renormalons \cite{renorm}: the Landau singularity in the perturbative expression for $\alpha_{\mbox{\tiny S}}(q^2)$ leads to a factorial divergence of perturbative expansions with respect to $\alpha_{\mbox{\tiny S}}(Q^2)$, where $Q^2$ is fixed and large, even if we remove the Landau singularity by making a non-perturbative modification of $\alpha_{\mbox{\tiny S}}(q^2)$ at low $q^2$ \cite{alfamodels}. In the case of eq.~(\ref{Kint}), the correction arising from the more careful treatment of the region $q_J^2<\mu_{\mbox{\tiny R}}^2$, corresponding to the second term in eq.~(\ref{Kom}), would be \begin{eqnarray}\label{Hintapp} \delta\int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_{j'}^2}{q_{j'}^2} \frac{d\phi_{j'}}{2\pi} K_\omega(q_J^2,\mu_{\mbox{\tiny R}}^2) &\simeq& \frac{\alb}{\om}\int_0^{\mu_{\mbox{\tiny R}}^2}\frac{dq_J^2}{q_j^2} \left[\left(\frac{\mu_{\mbox{\tiny R}}^2}{q_J^2}\right)^{\gamma_{\mbox{\tiny L}}(\alb/\om)} -\sum_{m=0}^\infty\left(-\frac{\alb}{\om} \ln\frac{q_J^2}{\mu_{\mbox{\tiny R}}^2}\right)^m\right]\nonumber\\ &=&\frac{\alb}{\om}\frac{\mu_{\mbox{\tiny R}}^2}{q_j^2} \left[\frac{1}{1-\gamma_{\mbox{\tiny L}}(\alb/\om)} -\sum_{m=0}^\infty m!\left(\frac{\alb}{\om}\right)^m\right]\;. \end{eqnarray} This does indeed contain a factorially divergent series, but, owing to the overall factor of $1/q_j^2$, it does not contribute any logarithms of $Q^2/\mu_{\mbox{\tiny R}}^2$ upon substitution in eq.~(\ref{G2def}). Therefore a more accurate treatment of the region $q_J^2<\mu_{\mbox{\tiny R}}^2$, although necessary to evaluate the integrals, does not affect the two-jet rate to SL precision. If we interpret the series in eq.~(\ref{G2SL}) as an asymptotic expansion, then the partial sum truncated after the smallest term represents an estimate of the total SL correction, with an uncertainty of the order of the smallest term. This estimate is shown by the points in figs.~\ref{fig_dR2a} and \ref{fig_dR2b}, with the uncertainty represented by the error bars. We see that estimates from the series are of the same order of magnitude as the numerical results, but the discrepancy may be several times the expected uncertainty. At the smaller value of $\bar\as/\omega$, the SL correction is relatively small (c.f.\ fig.~\ref{fig_R2a}) and the discrepancy is not so important. For the larger value of $\bar\as/\omega$, the estimate is better than expected, but the correction and the uncertainty are both large. \FIGURE{\input{corr2.pstex_t} \caption{\label{fig_dR2a} Solid: single-log correction to two-jet rate for $\bar\as/\omega = 0.2$. Points: estimate from asymptotic expansion; `error bars' indicate the smallest term in the expansion.}} \FIGURE{\input{corr4.pstex_t} \caption{\label{fig_dR2b} Solid: single-log correction to two-jet rate for $\bar\as/\omega = 0.4$. Points: estimate from asymptotic expansion; `error bars' indicate the smallest term in the expansion.}} To deduce the two-jet rate as a function of $x$, we can proceed as in eq.~(\ref{F1x}), writing \begin{equation}\label{F2x} F^{(2\,{\mbox{\scriptsize jet}})}(x,Q^2,\mu_{\mbox{\tiny R}}^2) = F(x,\mu_{\mbox{\tiny R}}^2)\otimes G^{(2)}(x,T) \end{equation} where \begin{equation}\label{G2x} G^{(2)}(x,T) = \frac{1}{2\pi i}\int_C d\omega\,x^{-\omega-1} \,G_\omega^{(2)}(T)\;. \end{equation} For the DL contribution we find from eq.~(\ref{G2DL}) that \begin{equation}\label{G2DLx} G^{(2,{\mbox{\scriptsize DL}})}(x,T) = \frac{1}{x\ln(1/x)}\,{\cal G}^{(2,{\mbox{\scriptsize DL}})}[\bar\as T\ln(1/x)] \end{equation} where \begin{equation}\label{cG2z} {\cal G}^{(2,{\mbox{\scriptsize DL}})}[z] = E_1(z)+\ln z +\gamma_{\mbox{\tiny E}}+1 +e^{-z}\left[E_1(-z)+\ln z +\gamma_{\mbox{\tiny E}}-1-z\right]\;, \end{equation} $E_1(z)$ being the exponential integral function \begin{equation}\label{E1z} E_1(z) = \int_z^\infty dt\frac{e^{-t}}{t}\;, \end{equation} interpreted as a principal-value integral when $z<0$. The divergence of the series for the SL correction in $\omega$-space, eq.~(\ref{G2SL}), is cured when one makes the inverse Mellin transformation to $x$-space, because the factorial coefficients are cancelled: \begin{equation}\label{melom} \frac{1}{2\pi i}\int_C d\omega\,x^{-\omega-1}\left(\frac{\alb}{\om}\right)^{2m+1} =\frac{\bar\as}{x}\frac{[\bar\as\ln(1/x)]^{2m}}{(2m)!}\;. \end{equation} Thus the SL correction to the two-jet rate can be expressed in closed form as a function of $x$: \begin{equation}\label{G2SLx} G^{(2,{\mbox{\scriptsize SL}})}(x,T) = \frac{1-x^{\bar\as T}}{x\ln(1/x)} \,{\cal G}^{(2,{\mbox{\scriptsize SL}})}[\bar\as\ln(1/x)] \end{equation} where \begin{equation}\label{cG2ans} {\cal G}^{(2,{\mbox{\scriptsize SL}})}[z] = \ln\Gamma(1-z)-\ln\Gamma(1+z)-2\gamma_{\mbox{\tiny E}} z\;. \end{equation} Notice that the expression (\ref{G2SLx}) is singular at $\bar\as\ln(1/x)=1$. From the viewpoint of the Mellin transformation (\ref{mellin}), it is this singularity that produces the divergence of the series in $\omega$-space. Conversely, use of the more correct expression (\ref{G2def}) in $\omega$-space, with the kernel function $K_\omega$ given in eq.~(\ref{Kom}), should suffice to remove the singularity in $x$-space. \subsection{Three-jet rate}\label{sec_3jet} The method used above for the two-jet rate can be extended, albeit laboriously, to higher jet multiplicities. In the three-jet case we have \begin{equation}\label{R3res} R^{(3\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2) = \frac{F^{(3\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2,\mu^2)}{F_\omega(Q^2,\mu^2)} \>=\>\exp[-\gamma_{\mbox{\tiny L}}(\alb/\om) T]\,G_\omega^{(3)}(T) \end{equation} where \begin{equation}\label{G3def} G_\omega^{(3)}(T) =\int_{\mu_{\mbox{\tiny R}}^2}^{Q^2}\frac{dq_j^2}{q_j^2} \frac{dq_{j'}^2}{q_{j'}^2}\frac{dq_{j''}^2}{q_{j''}^2} \frac{d\phi_{j'}}{2\pi}\frac{d\phi_{j''}}{2\pi} K_\omega(q_j^2,\mu_{\mbox{\tiny R}}^2) K_\omega(|q_j+q_{j'}|^2,\mu_{\mbox{\tiny R}}^2) K_\omega(|q_j+q_{j'}+q_{j''}|^2,\mu_{\mbox{\tiny R}}^2). \end{equation} One could in principle evaluate this expression numerically using eq.~(\ref{Kom}) for $K_\omega$. Here we derive the perturbative expansion analogous to eq.~(\ref{G2SL}). Introducing $t=\ln(q_j^2/\mu_{\mbox{\tiny R}}^2)$ etc.\ for brevity, the results in the appendix give \begin{equation} G_\omega^{(3)}(T) = \int_0^T dt\,H_\omega(t)\left[\int_0^T dt'\,L_\omega\left(\max\{t,t'\}\right) +2\sum_{m=1}^\infty\zeta(2m+1)\, \frac{\partial^{2m}L_\omega}{\partial t^{2m}}\right] \end{equation} where we write $H_\omega(q_j^2,\mu_{\mbox{\tiny R}}^2)$ as $H_\omega(t)$ and \begin{equation} L_\omega(t)= H_\omega(t)[\ln H_\omega(t) - \ln H_\omega(T)] + t\,[H_\omega(t)]^2 +2\sum_{m'=1}^\infty(2m')!\,\zeta(2m'+1)\,[H_\omega(t)]^{2m'+2}. \end{equation} Hence we obtain \begin{equation}\label{G3res} G_\omega^{(3)}(T)=G_\omega^{(3,{\mbox{\scriptsize DL}})}(T)+G_\omega^{(3,{\mbox{\scriptsize SL}})}(T) \end{equation} where $G_\omega^{(3,{\mbox{\scriptsize DL}})}$ is the double-logarithmic result \cite{Web98} \begin{eqnarray} &&G_\omega^{(3,{\mbox{\scriptsize DL}})}(T)\>=\> \frac{1}{6}\ln^3\left(1+\frac{\alb}{\om} T\right) +\ln^2\left(1+\frac{\alb}{\om} T\right)\nonumber\\ &&+\ln\left(1+\frac{\alb}{\om} T\right) \left(1+\frac{\alb}{\om} T\right)^{-1} -\frac{\alb}{\om} T\left(1+\frac{3\bar\as}{2\omega}T\right) \left(1+\frac{\alb}{\om} T\right)^{-2} \end{eqnarray} and \begin{eqnarray}\label{G3SL} &&G_\omega^{(3,{\mbox{\scriptsize SL}})}(T)\>=\> -2\sum_{m=1}^\infty (2m)!\,\zeta(2m+1) \left(\frac{\alb}{\om}\right)^{2m+1}\Biggl\{ \left[1-\left(1+\frac{\alb}{\om} T\right)^{-2m-2}\right]\nonumber\\ &&-\frac{1}{2m+1}\left[\ln\left(1+\frac{\alb}{\om} T\right) +\psi(2m+1)+\gamma_{\mbox{\tiny E}}+2\right]\left[1-\left(1+ \frac{\alb}{\om} T\right)^{-2m-1}\right]\Biggr\}+\nonumber\\ &&+\,4\sum_{m,m'=1}^\infty \frac{(2m+2m'+1)!}{(2m+2m'+2)(2m'+1)}\, \zeta(2m+1)\,\zeta(2m'+1) \left(\frac{\alb}{\om}\right)^{2m+2m'+2}\nonumber\\ &&\times \left[1-\left(1+\frac{\alb}{\om} T\right)^{-2m-2m'-2}\right]\;. \end{eqnarray} The expansion in eq.~(\ref{G3SL}) is again strongly divergent for all values of $\bar\as/\omega$. As discussed in subsect.~\ref{sec_2jet}, it can still be interpreted as an asymptotic expansion and used as a guide to the order of magnitude of the SL correction. Furthermore, the divergence is cured upon inverting the Mellin transformation to obtain the jet rate in $x$-space, as long as $\bar\as\ln(1/x)<1$. To extend the prediction to smaller values of $x$ one would need to evaluate eq.~(\ref{G3def}) numerically using the full kernel function in eq.~(\ref{Kom}). \section{Generating function for multi-jet rates}\label{sec_njet} For a general jet multiplicity $r$, we can write \begin{equation}\label{Rrres} R^{(r\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2) = \frac{1}{r!} \left.\frac{\partial^r}{\partial u^r}R_\omega(u,T)\right|_{u=0}\;, \end{equation} where the jet-rate generating function $R_\omega$ is given by \begin{equation} R_\omega(u,T)=\exp[-\gamma_{\mbox{\tiny L}}(\alb/\om) T]\,G_\omega(u,T) \end{equation} and \begin{equation}\label{GuT} G_\omega(u,T)=\sum_{r=0}^\infty u^r G^{(r)}_\omega(T)\;. \end{equation} The function $G^{(r)}_\omega(T)$ was given in eqs.~(\ref{G1def}), (\ref{G2res}) and (\ref{G3res}) for $r=1,2$ and 3, respectively. We can obtain the perturbative expansion of the function $G_\omega(u,T)$ as follows. We first define the unintegrated function $g_\omega(u,t,T)$ such that \begin{equation}\label{Gexpn} G_\omega(u,T)= 1+\int_0^T dt\,g_\omega(u,t,T)\;. \end{equation} Then using the results in the appendix we find that $g_\omega(u,t,T)$ satisfies the integro-differential equation \begin{equation}\label{gomu} g_\omega(u,t,T) = uH_\omega(t)\left[1+ \int_0^T dt'\,g_\omega\left(u,\max\{t,t'\},T\right) +2\sum_{m=1}^\infty\zeta(2m+1)\, \frac{\partial^{2m}g_\omega}{\partial t^{2m}}\right]\;. \end{equation} Writing \begin{equation} g_\omega(u,t,T) = \sum_{n=0}^\infty c_n(u,t,T) \left(\frac{\alb}{\om}\right)^n\;, \end{equation} this implies that \begin{eqnarray} c_{n+1}(u,t,T)&=&u\,\delta_{n,0}-(1-u)t\,c_n(u,t,T) +u\int_t^T dt'\,c_n(u,t',T)\nonumber\\ &&+2u\sum_{m=1}^\infty\zeta(2m+1) \frac{\partial^{2m}c_n}{\partial t^{2m}}\,. \end{eqnarray} Starting from $c_0=0$, this gives $c_n$ iteratively as a polynomial in $u$, $t$ and $T$, which can be substituted in eq.~(\ref{Gexpn}) to obtain the perturbative expansion of $G_\omega(u,T)$ to any desired order.\footnote{The results agree with those given to fourth order in ref.~\cite{M}.} The relation (\ref{melom}) can then be used to transform the result directly to $x$-space, giving \begin{equation}\label{Frresx} F^{(r\,{\mbox{\scriptsize jet}})}(x,Q^2,\mu_{\mbox{\tiny R}}^2) = \frac{1}{r!}F(x,\mu_{\mbox{\tiny R}}^2)\otimes \left.\frac{\partial^r}{\partial u^r}G(u,x,T)\right|_{u=0}\;, \end{equation} where \begin{equation}\label{GuxT} G(u,x,T)= \delta(1-x)+\int_0^T dt\,g(u,x,t,T) \end{equation} with \begin{equation} g(u,x,t,T) = \frac{\bar\as}{x}\sum_{n=0}^\infty c_{n+1}(u,t,T) \frac{[\bar\as\ln(1/x)]^n}{n!}\;, \end{equation} which we believe to be a convergent series as long as $\bar\as\ln(1/x)<1$. \subsection{Anomalous dimension} Notice that for $u=1$ we have \begin{equation} R_\omega(1,T)=\sum_{r=0}^\infty R^{(r\,{\mbox{\scriptsize jet}})}_\omega(Q^2,\mu_{\mbox{\tiny R}}^2) = 1 \end{equation} and therefore \begin{equation}\label{G1T} G_\omega(1,T)=\exp[\gamma_{\mbox{\tiny L}}(\alb/\om) T]\;. \end{equation} To show that eq.~(\ref{gomu}) does indeed lead to the Lipatov result (\ref{omlip}) for the anomalous dimension, we note that when $u=1$ the solution of eq.~(\ref{gomu}) is \begin{eqnarray}\label{gom1} g_\omega(1,t,T) &=& \gamma\,e^{\gamma(T-t)}\nonumber\\ G_\omega(1,T)&=& 1+\int_0^T dt\,g_\omega(1,t,T)=e^{\gamma T} \end{eqnarray} where \begin{eqnarray} \gamma &=& H_\omega(t)\left[1+\gamma t +2\sum_{m=1}^\infty\zeta(2m+1)\,\gamma^{2m+1}\right]\nonumber\\ &=& \frac{\alb}{\om}\left[1+ 2\sum_{m=1}^\infty\zeta(2m+1)\,\gamma^{2m+1}\right]\nonumber\\ &=& \frac{\alb}{\om}\gamma\left[\frac{1}{\gamma}-2\gamma_{\mbox{\tiny E}} -\psi(1+\gamma)-\psi(1-\gamma)\right]\;. \end{eqnarray} Rearranging terms, we obtain $\gamma=\gamma_{\mbox{\tiny L}}(\alb/\om)$ given by eqs.~(\ref{omlip}) and (\ref{glip}). \subsection{Jet multiplicity moments}\label{sec_mom} We can compute the moments of the jet multiplicity distribution by successively differentiating the generating function at $u=1$: \begin{equation} \langle r(r-1)\ldots(r-s+1)\rangle = \left.\frac{\partial^s}{\partial u^s}R_\omega(u,T)\right|_{u=1}\;. \end{equation} In this way we obtain the perturbative expansion of the mean number of jets \begin{eqnarray} \langle r\rangle &=& T\frac{\alb}{\om} +\frac{1}{2}T^2\left(\frac{\alb}{\om}\right)^2 +2\zeta(3)T\left(\frac{\alb}{\om}\right)^4\nonumber\\ &&+4\zeta(3)T^2\left(\frac{\alb}{\om}\right)^5 -8\zeta(5)T\left(\frac{\alb}{\om}\right)^6 +\cdots \end{eqnarray} and the mean square fluctuation in this number, \begin{eqnarray} \langle r^2\rangle -\langle r\rangle^2 &=& T\frac{\alb}{\om}+\frac{3}{2}T^2\left(\frac{\alb}{\om}\right)^2 +\frac{2}{3}T^3\left(\frac{\alb}{\om}\right)^3 -2\zeta(3)T\left(\frac{\alb}{\om}\right)^4\nonumber\\ &&+12\zeta(3)T^2\left(\frac{\alb}{\om}\right)^5 -\left(8\zeta(5)T -\frac{40}{3}\zeta(3)T^3\right) \left(\frac{\alb}{\om}\right)^6 +\cdots\;. \end{eqnarray} It appears true to all orders to SL precision that, as in the DL approximation \cite{Web98}, the mean number of jets is a quadratic function of $T$ and the mean square fluctuation is a cubic function of $T$. Thus the distribution of jet multiplicity at small $x$ and large $T$ is narrow, in the sense that its r.m.s.\ width increases less rapidly than its mean as $T$ increases. \section{Conclusions}\label{sec_conc} The calculation of jet rates at small $x$ poses many interesting challenges and sheds new light on the novel dynamics of this kinematic region. In the present paper we have concentrated on those perturbative contributions which have a factor of $\ln(1/x)$ for each power of $\alpha_{\mbox{\tiny S}}$ and are further enhanced by one or more powers of $T=\ln(Q^2/\mu_{\mbox{\tiny R}}^2)$, $\mu_{\mbox{\tiny R}}$ being the minimum resolved jet transverse momentum. For sufficiently large values of $\mu_{\mbox{\tiny R}}^2$ and $Q^2\gg\mu_{\mbox{\tiny R}}^2$, the resummation of such terms would seems to be a well-defined problem in perturbation theory. The results in sect.~\ref{sec_njet} do indeed specify all terms of the form $(\bar\as/\omega)^n T^m$ with $m>0$, for any jet multiplicity, $\omega$ being the moment variable in the Mellin transform. However, as we have seen explicity for the two- and three-jet rates (and we believe to be true more generally), the single-logarithmic terms (those with $0<m<n$) cannot be resummed directly since they form strongly divergent series. In $\omega$-space, the divergence is associated with kinematic regions in which the vector sums of transverse momenta of combinations of jets are less than $\mu_{\mbox{\tiny R}}$. A more careful treatment of such regions renders the jet rates well-defined as integrals. Furthermore, one obtains convergent series, within a limited range of $x$, after performing the inverse Mellin transformation to $x$-space. In the case of the two-jet rate, we were able to sum the resulting series explicitly, to obtain a closed-form expression valid in the region $\bar\as\ln(1/x)<1$. A number of interesting questions arise from our results. Clearly one would like to extend the resummation of jet rates to higher multiplicities and smaller values of $x$. This will require an $x$-space treatment of the difficult kinematic regions mentioned above. One would also like to prove the conjectures in subsect.~\ref{sec_mom} about jet multiplicity moments to all orders, and preferably to resum them. Ultimately, next-to-leading terms in $\ln(1/x)$ should also be included. Such terms will arise from next-to-leading corrections to the BFKL kernel and from the resolution of emitted gluons into two jets. \acknowledgments We thank G.\ Salam for valuable comments and especially for pointing out eq.~(\ref{HqltR}). BRW is also grateful to S.\ Catani and G.\ Marchesini for many helpful discussions.
\section{Introduction} \setcounter{equation}{0} \par More than twenty years ago many authors \cite{DSC} have proposed various strategies of deriving quark confinement in quantum chromodynamics (QCD). One of them is to show that the QCD vacuum is the dual superconductor which squeezes the color electric flux between quarks and anti-quarks. The evidences have been accumulated by recent investigations. Especially, recent numerical simulations have confirmed this picture, see \cite{review,Bali98}. In this scenario, the magnetic monopole \cite{Dirac31,WY75,GO78} obtained by the Abelian projection \cite{tHooft81} in QCD plays the essential role \cite{EI82,SY90}. These results suggest that the low-energy effective theory of QCD is given by the dual Ginzburg-Landau theory \cite{Suzuki88}. In fact, it has been shown that the dual Ginzburg-Landau theory can be derived starting from the QCD Lagrangian at least in the strong coupling region, see e.g, \cite{KondoI}. \par In the previous paper \cite{KondoII}, we have proposed a novel formulation of the Yang-Mills theory as a (perturbative) deformation of a topological quantum field theory (TQFT).% \footnote{In this reformulation, the gauge-fixed Yang-Mills theory is decomposed into the TQFT part and the remaining part. Then we assume that the remaining part can be treated in perturbation theory in the gauge coupling constant. This assumption is nothing but the meaning of the perturbative deformation. } We have shown \cite{KondoII,KondoIV} that the quark confinement in QCD in the sense of area law of the Wilson loop (or equivalently, the linear static potential between quark and anti-quark) can be derived from the formulation at least in the maximal Abelian (MA) gauge. The MA gauge realizes the Abelian magnetic monopole in Yang-Mills theory without introducing the scalar field as an elementary field (Hence it is realized as a composite field constructed from the gauge degrees of freedom). In the similar way, it has been shown \cite{KondoIII} that the four-dimensional Abelian gauge theory can have the confining phase in the strong coupling region. This can be used to give another derivation of quark confinement in QCD based on the low-energy effective {\it Abelian} gauge theory, see \cite{KondoV}. \par The above results are consistent with those of lattice gauge theory \cite{LGT}, though our formulation is given directly on the continuum space-time. This similarity is due to a fact that the ingredients of confinement in our formulation lies in the compactness of the gauge group (or the periodicity in the gauge potential) and the existence of topological soliton. Thus, the existence of magnetic monopole is a sufficient condition for explaining quark confinement, as confirmed by analytical and numerical results \cite{review}. \par In this paper, we re-derive the formulation proposed in \cite{KondoII,KondoIII} based on the background field method (BGFM) \cite{DeWitt67,tHooft75,Abbott82,Abbott81,AGS83}. A purpose of this paper is to fill the gap in the previous presentation \cite{KondoII} without any ad hoc argument. This derivation enables us to discuss various topological soliton or topological defect other than the magnetic monopole, which might equally play the important role in explaining the origin of quark confinement. Such a viewpoint is necessary to answer the question: what are the most relevant degrees of freedom for quark confinement, since the necessary and sufficient condition for quark confinement is not yet known. Therefore, our formulation can also be applied to other scenarios of quark confinement based on various confiners, e.g. instanton, center vortex or non-Abelian magnetic monopole, although the details will be given in a subsequent paper. Another advantage of BGFM is that it simplifies the proof \cite{QR97,KondoI} that the renormalization group beta function of the Abelian-projected effective gauge theory is the same as the original Yang-Mills non-Abelian gauge theory. \par Just as the topological Yang-Mills theory \cite{Witten88} describes the gauge field configurations satisfying the self-dual equation, i.e., instantons, \begin{eqnarray} {\cal F}_{\mu\nu}=\pm \tilde {\cal F}_{\mu\nu} , \quad \tilde {\cal F}_{\mu\nu} := {1 \over 2} \epsilon_{\mu\nu\rho\sigma} {\cal F}^{\rho\sigma} , \label{SDeq0} \end{eqnarray} the TQFT that we have proposed deals with the gauge field configurations which obeys the MA gauge equation, \begin{eqnarray} D^{\mp}_\mu[a]A_\mu^{\pm}=0 , \label{MAGeq} \end{eqnarray} which is nothing but the background field equation. Both equations are the 1st order partial differential equations. They may have some properties in common. In fact, a class of classical solutions of the MA gauge equation (\ref{MAGeq}) simultaneously satisfies the self-dual equation (\ref{SDeq0}) and vice versa \cite{BOT96,CG95}. It is obtained from the same ansatz as that of 't Hooft for the multi-instanton. The instanton is the point defect in four dimensions, while the magnetic monopole is the point defect in three dimensions. In four dimensions, therefore, the magnetic monopole is a one-dimensional object, i.e., a current $k_\mu$ (a closed loop due to the topological conservation law $\partial_\mu k_\mu = 0$). This is a Lorentz covariant generalization of the observation that the static monopole in three dimensions draws the straight line in the time direction in four dimensions where the monopole charge is given by the integral $Q_m:=\int d^3x k_0(x)$ from the monopole density $k_0$. Since we frequently use the 'magnetic' monopole as implying the solution of the MA gauge equation, the solution can describe the object which looks like the magnetic monopole and the instanton at the same time. Therefore, the magnetic monopole and the instanton are not the disjoint concept in four dimensions. Actually, strong correlations between monopoles and instantons are shown in the analytical studies \cite{BOT96,CG95} and observed in the lattice simulations \cite{HT96,BS96,STSM95,FSST97,FMT97}. It is easy to see that the instanton is also a solution of the field equation (2nd order partial differential equation) \begin{eqnarray} {\cal D}_\nu[{\cal A}] {\cal F}_{\mu\nu}=0 . \label{Feq0} \end{eqnarray} However, it is not yet clarified which solution of the MA gauge equation (\ref{MAGeq}) becomes that of the field equation besides the solution mentioned above. The solution of (\ref{MAGeq}) may contain the solution which is not the solution of the field equation (\ref{Feq0}). \par When we see the intersection of the magnetic monopole current with the two-dimensional plane, the classical configuration satisfying (\ref{MAGeq}) looks like the instanton in two-dimensional nonlinear sigma model (NLSM$_2$), as shown in \cite{KondoII}. Therefore the condensation of the magnetic monopole current in four dimensions can be examined on the two-dimensional subspace which can be chosen arbitrarily. The condensation of the two-dimensional instanton in NLSM$_2$ leads to that of the four-dimensional magnetic monopole current in Yang-Mills theory. So the instanton condensation in NLSM$_2$ is a sufficient condition of quark confinement based on the dual superconductor scenario. \par In the scenario \cite{KondoII} of deriving quark confinement, the gauge fixing part for the gauge fixing condition (\ref{MAGeq}) has played the essential role. Since the quark confinement must be a gauge invariant concept, it is better to derive it based on the gauge invariant formulation. In contrast to the lattice gauge theory, however, the continuum formulation of the ordinary gauge theory free from the gauge fixing is not available except for special cases. Then we are forced to deal with the formulation based on the specific choice of gauge fixing. The readers might think the claim strange that the essence of quark confinement lies in the gauge-fixing part. Recall that, in the level of the classical theory, the action part of the gauge theory is well understood from a viewpoint of the geometry of connection. In quantum theory, however, we need to include the gauge fixing term in order to correctly quantize the gauge theory. Usually, the gauge fixing term introduced in this way is not considered to have any geometric meaning. However, this observation is not necessarily correct. In fact, the gauge fixing term plus the associated Faddeev-Popov ghost term can have the very geometric meaning from the viewpoint of global topology, as will be discussed in this paper. In the quantum gauge theory, therefore, the action part and the gauge fixing part should be treated on equal footing. Unfortunately, we must discuss the topology of the infinite dimensional manifold for gauge field configurations. Then the mathematically rigorous analysis will be rather hard, so that we can at best analyze the finite dimensional analog. \par Usually, we consider that, even if the gauge fixing term has a geometric interpretation, it can not have any local dynamics (propagating mode) and describe only the topological objects, since it is written as the Becchi-Rouet-Stora-Tyupin (BRST) exact form, i.e., $S_{GF}=\{ Q_B, \kappa \Psi \}$ using the BRST charge $Q_B$. In the manifestly covariant formalism of gauge theory, the physical state $| phys \rangle$ is specified by the condition, $Q_B | phys \rangle =0$. If we consider the theory with the action $S_{GF}=\{ Q_B, \kappa \Psi \}$ alone by neglecting the Yang-Mills action (this theory is identified with the TQFT), the expectation value of the gauge invariant quantity $\langle {\cal O} \rangle$ does not depend on the coupling $\kappa$, since ${\partial \over \partial \kappa} \langle 0|{\cal O} |0 \rangle = - \langle 0| {\cal O} \{ Q_B, \Psi \} |0 \rangle + \langle 0| {\cal O} |0 \rangle \langle 0|\{ Q_B, \Psi \} |0 \rangle = - \langle 0| \{ Q_B, {\cal O} \Psi \} \rangle = 0 $ where we have used the BRST invariance of ${\cal O}$ and $|0 \rangle \in |phys \rangle$. However, taking into account the action $S_{YM}$ in addition to $S_{GF}$, we can not draw the same conclusion. This is the usual situation of quantized gauge theory. A subtle point is that the above consideration is based on the assumption that the BRST symmetry is not broken. If the BRST symmetry happen to be spontaneously broken \cite{Fujikawa83}, the physical state including the vacuum is not annihilated by the BRST charge, i.e., $Q_B | phys \rangle \not=0$. In this case, the TQFT with the action $S_{TQFT}=\{ Q_B, \kappa \Psi \}$ can have local dynamics and the expectation value can depend on the coupling constant $\kappa$. \par In our scenarios, the spontaneous breaking of the hidden supersymmetry $OSp(4|2)$ rather than the BRST symmetry can take place by the dimensional reduction (in the sense of Parisi-Sourlas \cite{PS79,HK85}), at least for a special choice of the MA gauge \cite{KondoII}. Here it is worth remarking that the equivalence of the correlation functions hold only for a class of them and hence the Hilbert space of the reduced theory is different from the original theory \cite{KondoII}. The symmetry breaking occurs spontaneously in the following sense. We can choose arbitrary $(D-2)$-dimensional subspace from the $D$-dimensional spacetime. Once the specific subspace is chosen, however, the hidden supersymmetry $OSp(4|2)$, i.e., the rotational symmetry in the superspace is broken by this procedure. \par Another purpose of this paper is to propose a numerical simulation in order to confirm the dimensional reduction and examine its implications to quark confinement problem. The result will prove or disprove the validity of our scenario for deriving quark confinement based on the above reformulation. \par This paper is organized as follows. In section 2, we briefly review the BGFM for the Yang-Mills theory and its BRST version based on the functional integral formalism. In section 3, we explain how the quantum theory of topological soliton can be obtained in the framework of BGFM. We discuss a relationship between the instanton and the magnetic monopole in this construction. In section 4, by making the change of gauge field variable, we show that the formulation proposed in \cite{KondoII} is recovered from the BFGM. This is the main result of this paper. In section 5, we give a strategy of deriving quark confinement based on the above formulation. We take up some issues which have not been mentioned in the previous publications. We give a proposal of numerical calculation for checking the validity of the strategy. In section 6, we examine the mass generation for the gluon field in the MA gauge. We discuss a possibility of mass generation caused by the dimensional reduction as a result of breakdown of the hidden supersymmetry. In section 7, we discuss that the gauge fixing part in the quantum theory of gauge fields can have a geometric meaning from the viewpoint of global topology. In the final section, we summarize the results and discuss the role of various topological solitons other than the magnetic monopole for explaining color confinement in QCD. \section{Background field method} \setcounter{equation}{0} \par \subsection{Path integral for Yang-Mills field} We consider the functional integral approach to the Yang-Mills gauge field theory with the action \begin{eqnarray} S_{YM}[{\cal A}] := \int d^Dx {\cal L}_{YM}[{\cal A}] = - \int d^Dx {1 \over 4} ({\cal F}_{\mu\nu}^A[{\cal A}])^2 , \end{eqnarray} where ${\cal F}_{\mu\nu}^A[{\cal A}]$ is the field strength for the gauge field ${\cal A}_\mu^A$ defined by \begin{eqnarray} {\cal F}_{\mu\nu}^A[{\cal A}] := \partial_\mu {\cal A}_\nu^A - \partial_\nu {\cal A}_\mu^A + g f^{ABC} {\cal A}_\mu^B {\cal A}_\nu^C . \end{eqnarray} \par In the quantum theory of the Yang-Mills gauge field, the generating functional is defined by \footnote{The tilde is used only for later convenience (in section 4) and does not have particular physical meaning. } \begin{eqnarray} Z[J] := \int [d{\cal A}] \delta(\tilde F^A[{\cal A}]) \det \left[ {\delta \tilde F^A \over \delta \tilde \omega^B} \right] \exp \left\{ i [ S_{YM}[{\cal A}] + (J_\mu \cdot {\cal A}_\mu) ] \right\} , \label{Z[J]} \end{eqnarray} where $(J \cdot {\cal A})$ is the source term \begin{eqnarray} (J_\mu \cdot {\cal A}_\mu):= \int d^Dx J_\mu^A(x) {\cal A}_\mu^A(x) . \end{eqnarray} In (\ref{Z[J]}), the gauge-fixing condition is imposed by \begin{eqnarray} \tilde F^A[{\cal A}] = 0 , \end{eqnarray} and $ \det \left[ {\delta \tilde F^A \over \delta \tilde \omega^B} \right] $ is the so-called Faddeev-Popov (FP) determinant which is the determinant of the derivative of the gauge-fixing function $\tilde F^A$ under an infinitesimal gauge transformation, \begin{eqnarray} \delta {\cal A}_\mu^A = {\cal D}_\mu^{AB}[{\cal A}] \tilde \omega^B := \partial_\mu \tilde \omega^A + gf^{ABC} {\cal A}_\mu^B \tilde \omega^C , \\ {\cal D}_\mu^{AB} := \partial_\mu \delta^{AB} - gf^{ABC} {\cal A}_\mu^C . \end{eqnarray} The delta function is made less singular by introducing the gauge-fixing parameter $\tilde \alpha$ as \begin{eqnarray} \delta(\tilde F^A[{\cal A}]) := \prod_{x, A} \delta(\tilde F^A[{\cal A}(x)]) \rightarrow \exp \left\{ -i{1 \over 2\tilde \alpha} (\tilde F[{\cal A}] \cdot \tilde F[{\cal A}]) \right\} . \end{eqnarray} \par For example, a common choice is the Lorentz gauge, \begin{eqnarray} \tilde F^A[{\cal A}] = \partial_\mu {\cal A}_\mu^A . \end{eqnarray} Then the FP determinant is given by \begin{eqnarray} \det \left[ {\delta \tilde F^A \over \delta \tilde \omega^B} \right] = \det (\partial_\mu {\cal D}_\mu^{AB}[{\cal A}]\delta^D(x-y)) . \end{eqnarray} The connected Green's functions are generated by \begin{eqnarray} W[J] := - i \ln Z[J] . \end{eqnarray} The effective action is defined by making the Legendre transformation \begin{eqnarray} \Gamma[\bar Q] := W[J] - (J_\mu \cdot \bar Q_\mu) , \end{eqnarray} where \begin{eqnarray} \bar Q^A := {\delta W \over \delta J_\mu^A} . \end{eqnarray} It is well known that the derivative of the effective action with respect to $\bar Q$ are the one-particle irreducible (1PI) Green's function. \subsection{BGFM} \par Next, we consider the quantization on a given background gauge field $\Omega_\mu$, \begin{eqnarray} {\cal A}_\mu = \Omega_\mu + {\cal Q}_\mu , \label{separa} \end{eqnarray} where ${\cal Q}_\mu$ denotes the field to be quantized. The generating functional is given by \begin{eqnarray} \tilde Z[J, \Omega] := \int [d{\cal Q}] \det \left[ {\delta \tilde F^A \over \delta \tilde \omega^B} \right] \exp \left\{ i \left[ S_{YM}[\Omega+{\cal Q}] + (J_\mu \cdot {\cal Q}_\mu) -{1 \over 2\tilde \alpha} (\tilde F[{\cal Q}] \cdot \tilde F[{\cal Q}]) \right] \right\} . \label{Z[J,O]} \end{eqnarray} where the gauge invariance for ${\cal Q}_\mu$ is broken by the the gauge fixing condition $\tilde F^A[{\cal Q}]=0$ which is supposed to fix completely the gauge degrees of freedom and the ${\delta \tilde F^A \over \delta \tilde \omega^B}$ is the derivative of the gauge-fixing term under the infinitesimal gauge transformation given by \begin{eqnarray} \delta {\cal Q}_\mu^A = ({\cal D}_\mu[\Omega+{\cal Q}] \tilde \omega)^A . \end{eqnarray} In (\ref{Z[J,O]}), we do not couple the background field to the source following 't Hooft \cite{tHooft75} In the background field method (BGFM) \cite{DeWitt67,tHooft75,Abbott82,Abbott81,AGS83}, the the following gauge fixing condition is chosen, \begin{eqnarray} \tilde F^A[{\cal Q}] := {\cal D}_\mu^{AB}[\Omega] {\cal Q}_\mu^B = 0 , \label{BGFgauge} \end{eqnarray} which is called the background field (BGF) gauge. An advantage of the BGF gauge is that the BGF gauge condition retains explicit gauge invariance for the background gauge field $\Omega_\mu$ even after the gauge fixing for the field ${\cal Q}_\mu$. \par {\it Proposition}\cite{Abbott82}: Under the BGF gauge condition (\ref{BGFgauge}), the BGF generating functional $\tilde Z[J, \Omega]$ and $\tilde W[J, \Omega] := -i \ln \tilde Z[J, \Omega]$ are invariant under the (infinitesimal) transformation, \begin{eqnarray} \delta \Omega_\mu^A &=& ({\cal D}_\mu[\Omega] \omega)^A := (\partial_\mu \omega + i g [\omega, \Omega_\mu])^A , \label{traOme} \\ \delta J_\mu^A &=& i g[\omega, J_\mu]^A := - g f^{ABC} \omega^B J_\mu^C . \label{traJ} \end{eqnarray} \par This is shown as follows. By making the change of integration variables, ${\cal Q}_\mu \rightarrow {\cal Q}_\mu + i[\omega, {\cal Q}_\mu]$, i.e., \begin{eqnarray} \delta {\cal Q}_\mu^A = i g[\omega, {\cal Q}_\mu]^A = \omega \times {\cal Q} . \label{traQ} \end{eqnarray} Eq.~(\ref{traJ}) and (\ref{traQ}) represent an adjoint group rotation for $J_\mu$ and ${\cal Q}_\mu$ respectively, so the term $(J_\mu \cdot Q_\mu)$ is clearly invariant. Adding (\ref{traOme}) and (\ref{traQ}), we find \begin{eqnarray} \delta (\Omega_\mu+{\cal Q}_\mu)^A = ({\cal D}_\mu[\Omega+{\cal Q}] \omega)^A . \label{traA} \end{eqnarray} This is just a gauge transformation on the field variable ${\cal A}_\mu=\Omega_\mu+{\cal Q}_\mu$, so the action $S_{YM}[\Omega+{\cal Q}]$ is also invariant. Note that the BGF gauge condition $\tilde F^A[{\cal Q}]$ is just the covariant derivative of ${\cal Q}_\mu$ with respect to the BGF $\Omega_\mu$. Eq.~(\ref{traOme}) is a gauge transformation on $\Omega_\mu$ and (\ref{traQ}) is an adjoint rotation of ${\cal Q}_\mu$. Then the gauge fixing term $(\tilde F \cdot \tilde F)$ is invariant under such transformations. The FP determinant is also invariant, since the determinant is invariant under the adjoint rotation. Thus the BGF generating functional $\tilde Z[J, {\cal Q}]$ is invariant under (\ref{traOme}) and (\ref{traJ}). \par By using \begin{eqnarray} \tilde W[J, \Omega] := -i \ln \tilde Z[J, \Omega] , \end{eqnarray} we define the background effective action \begin{eqnarray} \tilde \Gamma[\tilde {\cal Q}, \Omega] := \tilde W[J, \Omega] - (J_\mu, \tilde Q_\mu) , \end{eqnarray} where \begin{eqnarray} \tilde Q_\mu^A = {\delta \tilde W \over \delta J_\mu^A} . \end{eqnarray} From the invariance of $\tilde Z[J, {\cal Q}]$, it follows that $\tilde \Gamma[\tilde {\cal Q}, \Omega]$ is invariant under \begin{eqnarray} \delta \Omega_\mu &=& ({\cal D}_\mu[\Omega] \tilde \omega) , \label{traOme2} \\ \delta \tilde {\cal Q}_\mu &=& i g[\tilde \omega, \tilde {\cal Q}_\mu] , \label{traQ2} \end{eqnarray} Since (\ref{traQ2}) is a homogeneous transformation, $\tilde \Gamma[0, \Omega]$ is invariant under the transformation (\ref{traOme2}) alone. Hence the effective action $\tilde \Gamma[0, \Omega]$ in the BGFM is an explicitly gauge invariant functional of $\Omega$, since (\ref{traOme2}) is just an ordinary gauge transformation. As a result, 1PI Green's functions generated by differentiating $\tilde \Gamma[0, \Omega]$ with respect to $\Omega$ will obey the naive Ward-Takahashi identities of gauge invariance. Hence, $\tilde \Gamma[0, \Omega]$ calculated in the BGFG is equal to the conventional effective action $\Gamma[\bar Q]$ with $\bar Q=\Omega$ calculated in an unconventional gauge which depends on $\Omega$ \begin{eqnarray} \tilde F^A[{\cal Q}] := {\cal D}_\mu^{AB}[\Omega] ({\cal Q}_\mu^B - \Omega_\mu^B) = \partial_\mu {\cal Q}_\mu^A + g f^{ABC} \Omega_\mu^B {\cal Q}_\mu^C - \partial_\mu \Omega_\mu = 0 . \label{BGFgauge2} \end{eqnarray} Then we obtain \begin{eqnarray} \tilde \Gamma[0, \Omega] = \Gamma[\bar Q]\big|_{\bar Q=\Omega} , \end{eqnarray} as a special case of \begin{eqnarray} \tilde \Gamma[\tilde Q, \Omega] = \Gamma[\bar Q]\big|_{\bar Q = \tilde Q + \Omega} . \end{eqnarray} The 1PI Green functions calculated from the gauge invariant effective action $\tilde \Gamma[0,\Omega]$ will be very different from those calculated by conventional method in normal gauges. Nevertheless, the relation assures us that all gauge-invariant physical quantities will come out the same in either approach \cite{AGS83}. Thus $\tilde \Gamma[0, \Omega]$ can be used to generate the S-matrix of a gauge theory in exactly the same way as the usual effective action is employed. \subsection{BRST version of the BGFM} \par Now we give the Becchi-Rouet-Stora-Tyupin (BRST) version of the BGFM. The BGF generating functional is rewritten into \begin{eqnarray} \tilde Z[J, \Omega] := \int [d{\cal Q}] [d\tilde C][d\bar {\tilde C}][d\tilde B] \exp \left\{ i S_{YM}[\Omega+{\cal Q}] + i \tilde S_{GF}[{\cal Q}, \tilde C, \bar {\tilde C}, \tilde B] + i (J_\mu \cdot {\cal Q}_\mu) \right\} , \label{ZBGFM} \end{eqnarray} where $\tilde B$ is the auxiliary scalar field and $\tilde C, \bar {\tilde C}$ are Hermitian anticommuting scalar field called the FP ghost and anti-ghost field, $\tilde C^\dagger = \tilde C, \bar {\tilde C}^\dagger=\bar {\tilde C}$. Using the BRST transformation, \begin{eqnarray} \tilde \delta_B \Omega_\mu(x) &=& 0 , \nonumber\\ \tilde \delta_B {\cal Q}_\mu(x) &=& {\cal D}_\mu[\Omega+{\cal Q}] {\tilde C}(x) := \partial_\mu {\tilde C}(x) - ig [\Omega_\mu(x)+{\cal Q}_\mu(x), {\tilde C}(x)], \nonumber\\ \tilde \delta_B {\tilde C}(x) &=& i{1 \over 2}g[{\tilde C}(x), {\tilde C}(x)], \nonumber\\ \tilde \delta_B \bar {\tilde C}(x) &=& i \tilde B(x) , \nonumber\\ \tilde \delta_B \tilde B(x) &=& 0 , \label{BRST0} \end{eqnarray} the gauge fixing and the FP ghost terms for the BGG are combined into a compact form, \begin{eqnarray} \tilde S_{GF}[{\cal Q}, \tilde C, \bar {\tilde C}, \tilde B] &:=& - \int d^Dx \ i \tilde \delta_B \ {\rm tr}_G \left[ \bar {\tilde C}\left( \tilde F[{\cal Q}]+ {\tilde \alpha \over 2}\tilde B \right) \right] , \label{GF0} \end{eqnarray} or \begin{eqnarray} \tilde S_{GF}[{\cal Q}, \tilde C, \bar {\tilde C}, \tilde B] = \int d^Dx \ {\rm tr}_G \left[ \tilde B {\cal D}_\mu[\Omega] {\cal Q}_\mu + {\tilde \alpha \over 2} \tilde B \tilde B + i \bar {\tilde C} {\cal D}_\mu[\Omega] {\cal D}_\mu[\Omega+{\cal Q}]\tilde C \right] , \end{eqnarray} where $\tilde \alpha$ is the gauge-fixing parameter and $\tilde \alpha=0$ corresponds to the Landau gauge (delta function gauge). This is clearly BRST invariant $\delta_B S_{GF}= 0$ due to nilpotency of the BRST transformation, $\delta_B^2 \equiv 0$. If the auxiliary field $\tilde B$ is integrated out, the gauge-fixing part reads \begin{eqnarray} \tilde S_{GF}[{\cal Q}, \tilde C, \bar {\tilde C}] = \int d^Dx \ {\rm tr}_G \left[ -{1 \over 2\tilde \alpha} ({\cal D}_\mu[\Omega] {\cal Q}_\mu)^2 + i \bar {\tilde C} {\cal D}_\mu[\Omega] {\cal D}_\mu[\Omega+{\cal Q}]\tilde C \right] . \label{GFQ} \end{eqnarray} In fact, this recovers the original form (\ref{Z[J,O]}), since \begin{eqnarray} \det \left[ {\delta \tilde F^A \over \delta \tilde \omega^B} \right] = \int [d\tilde C][d\bar {\tilde C}] \exp \left[ i\int d^Dx \ {\rm tr}_G \left( i \bar {\tilde C} {\cal D}_\mu[\Omega] {\cal D}_\mu[\Omega+{\cal Q}]\tilde C \right) \right] , \end{eqnarray} The explicit form of the FP ghost term is \begin{eqnarray} {\rm tr}_G \left[ i \bar {\tilde C} {\cal D}_\mu[\Omega] {\cal D}_\mu[\Omega+{\cal Q}]\tilde C \right] &=& i \bar {\tilde C}^A [ \partial_\mu \partial_\mu \delta^{AB} - g f^{ACB} {\uparrow \partial_\mu} (\Omega_\mu+{\cal Q}_\mu)^C + g f^{ACB} \Omega_\mu^B \partial_\mu \nonumber\\&& + g^2 f^{ACE}f^{EDB} \Omega_\mu^C (\Omega_\mu+{\cal Q}_\mu)^D ] \tilde C^B , \label{GFQ1} \end{eqnarray} and the gauge fixing term is \begin{eqnarray} && {\rm tr}_G \left[ -{1 \over 2\tilde \alpha} ({\cal D}_\mu[\Omega] {\cal Q}_\mu)^2 \right] \nonumber\\ &=& -{1 \over 2\tilde \alpha} \left[ (\partial_\mu {\cal Q}_\mu^A)^2 + 2 g f^{ABC} \Omega_\nu^B {\cal Q}_\nu^C \partial_\mu {\cal Q}_\mu^A + g^2 f^{ABC} f^{ADE} \Omega_\mu^B {\cal Q}_\mu^C \Omega_\nu^D {\cal Q}_\nu^E \right] . \label{GFQ2} \end{eqnarray} Feynmann rule for the BGFM is derived from the shifted action $S_{YM}[\Omega+{\cal Q}]$ and (\ref{GFQ}), see Abbott \cite{Abbott82}. In the limit $\Omega_\mu \rightarrow 0$, the BRST version of BGFM reduces to the usual BRST formulation of the Yang-Mills theory in the Lorentz gauge, $F^A[Q] = \partial^\mu {\cal Q}_\mu$. \par The advantage of the BGFM becomes apparent when the two-loop $\beta$ function is calculated. The BGFM makes the calculation much easier than previous calculations using the conventional approach, see \cite{Abbott81,AGS83}. \section{Quantum theory of topological soliton and BGFM} \setcounter{equation}{0} \subsection{Summation over topological soliton background} \par In the conventional approach, the background field $\Omega_\mu$ is chosen to be a solution of the classical field equation. In Yang-Mills theory, the equation of motion is given by \begin{eqnarray} {\delta S_{YM}[{\cal A}] \over \delta {\cal A}_\mu^A} \equiv {\cal D}^{AB}_{\nu} [{\cal A}] {\cal F}_{\mu\nu}^{B}[{\cal A}] = 0. \label{Feq} \end{eqnarray} Then, under the identification (\ref{separa}), \begin{eqnarray} {\cal A}_\mu = \Omega_\mu + {\cal Q}_\mu , \end{eqnarray} the quantization is performed around arbitrary but fixed background $\Omega_\mu$ which satisfies (\ref{Feq}). In this paper, we consider the topologically nontrivial field configuration as a background field $\Omega_\mu$, around which the quantization of the Yang-Mills theory is performed. Once a specific type of field configurations is chosen as the background $\Omega_\mu$, we will include all possible configurations of the same type, in other words, we sum up all contributions coming from such a type of configurations. \footnote{ Such a procedure was performed so far in various forms, e.g., by summing up the monopole-currents trajectories \cite{ST78,BS78,AE99}. } Therefore, in our formulation, a candidate for the generating functional of the {\it total} Yang-Mills theory is given by \begin{eqnarray} Z[J] &=& \int [d\Omega_\mu] \tilde Z[J, \Omega] =: \int [d\Omega_\mu] \exp (i \tilde S_{eff}[J, \Omega]) , \label{Z[J]0} \end{eqnarray} where we have defined \begin{eqnarray} \tilde S_{eff}[J, \Omega] := - i \ln \tilde Z[J, \Omega] , \end{eqnarray} and $[d\Omega_\mu]$ is the integration measure specified later. \par Note that the action $\tilde S_{eff}[J, \Omega]$ can have the local gauge invariance (\ref{traOme}) for $\Omega_\mu$ by virtue of the BGFM. Hence, the total Yang-Mills theory defined in this way is identified with the (quantized) gauge theory with the action $\tilde S_{eff}[J,\Omega]$, provided that the integration measure $[d\Omega_\mu]$ is gauge invariant. However, in order to quantize the total Yang-Mills theory correctly, we need to fix the local gauge invariance for the non-Abelian gauge field $\Omega_\mu$. Thus, instead of (\ref{Z[J]0}), we define the generating functional of the total Yang-Mills theory by \begin{eqnarray} Z[J] = \int [d\Omega_\mu] \delta(F^A[\Omega]) \det \left[ {\delta F^A \over \delta \omega^B} \right] \tilde Z[J, \Omega] , \end{eqnarray} or \begin{eqnarray} Z[J] = \int [d\Omega_\mu] \det \left[ {\delta F^A \over \delta \omega^B} \right] \exp \left(i \tilde S_{eff}[J, \Omega] -i{1 \over 2\alpha} (F[\Omega] \cdot F[\Omega]) \right) , \end{eqnarray} where the gauge fixing function $F^A$ is not necessarily equal to the BGF gauge $\tilde F^A$. The choice of $F^A$ is quite important in our formulation for realizing topological soliton background, as explained below. In order to be able to incorporate the topological soliton, the gauge fixing function $F[\Omega]$ should be {\it nonlinear} in $\Omega$. The measure $[d\Omega_\mu]$ must be chosen appropriately for the topological soliton in question. In the final stage the measure is replaced by the integration over the collective coordinates of the soliton. \begin{figure} \begin{center} \unitlength=1cm \begin{picture}(12,8) \thicklines \put(3,7.5){\framebox(6,1){Yang-Mills classical solution}} \put(4,7.0){2nd order NL PDE} \put(4,6.5){${\cal D}_\nu[{\cal A}] {\cal F}_{\mu\nu}=0$} \put(6,3.5){\oval(13,8)} \put(8.5,7){\circle{0.2}} \put(8.8,6.9){SSU} \put(1.5,5){\framebox(4,1){Magnetic monopole}} \put(1.5,4.5){1st order NL PDE} \put(1.5,4.0){MA gauge} \put(1.5,3.5){$D^{\mp}_\mu[a]A_\mu^{\pm}=0$} \put(-1.0,2){\bf $?$} \put(3.0,2.5){\oval(9,5)} \put(5.5,2){\circle*{0.2}} \put(5.8,1.9){'t Hooft} \put(1.5,2){\circle{0.2}} \put(1.8,0.9){CG} \put(1.5,1){\circle{0.2}} \put(1.8,1.9){BOT} \put(8,4.5){\framebox(3,1){Instanton}} \put(8,4.0){Ist order NL PDE} \put(8,3.5){${\cal F}_{\mu\nu}=\pm \tilde {\cal F}_{\mu\nu}$} \put(8,3.0){ADHM} \put(8.5,2.5){\oval(7,4)} \put(8.0,1){\circle{0.2}} \put(8.3,0.9){Witten} \end{picture} \end{center} \caption{ Moduli space, i.e, space of solutions for the Yang-Mills equation of motion (\ref{Feq}), self-dual instanton equation (\ref{SDeq}) and the magnetic monopole equation (\ref{MAg0}) in the MA gauge in four dimensions. The instanton solution of the self-dual equation (\ref{SDeq}) is also a solution of the Yang-Mills equation of motion (\ref{Feq}). The converse is not necessarily true. In fact, the Sibner-Sibner-Uhlenbeck (SSU) solution \cite{SSU89} on $S^4$ is a solution of the Yang-Mills field equation which is not a solution of the self-dual equation \cite{SSU89}. The general instanton solution on $S^4$ can be constructed according to the Atiyah, Drinfeld, Hitchin and Mannin (ADHM) \cite{ADHM}. The explicit form for the multi-instanton is known in the specific cases, e.g., 't Hooft type \cite{Rajaraman89} or Witten type \cite{Witten77}. Both types include one-instanton solution of Belavin, Polyakov, Schwartz and Tyupkin (BPST) \cite{BPST75}. The multi-instanton solution of 't Hooft type is also the solution of the magnetic monopole equation (\ref{MAg}). Some solutions are known for (\ref{MAg}), Chernodub and Gubarev (CG) \cite{CG95} and Brower, Orginos and Tan (BOT) \cite{BOT96}. See Appendix A. The general solution of (\ref{MAg}) is not yet known. In principle, there may exist a solution of the monopole equation which is not a solution of Yang-Mills field equation, indicated by $?$ in the figure. } \label{fig:moduli} \end{figure} \subsection{Yang-Mills instanton} \par In four-dimensional Euclidean space, the most popular topologically nontrivial field configuration of pure Yang-Mills theory is the instanton (anti-instanton) \cite{BPST75,tHooft76,Witten77,JNR77,ADHM,Coleman85,Rajaraman89} which is a solution of the self-dual (self-antidual) equation, \begin{eqnarray} {\cal F}_{\mu\nu}[{\cal A}] = \pm {\cal F}_{\mu\nu}^*[{\cal A}], \quad {\cal F}_{\mu\nu}^*[{\cal A}] := {1 \over 2} \epsilon_{\mu\nu\rho\sigma} {\cal F}_{\rho\sigma}[{\cal A}] , \label{SDeq} \end{eqnarray} with a finite action \begin{eqnarray} S_{YM}[{\cal A}] < \infty . \end{eqnarray} The self-dual equation (\ref{SDeq}) is a first order nonlinear partial differential equation (NL PDE), whereas the field equation (\ref{Feq})is a second order nonlinear partial differential equation. The instanton is a kind of topological soliton which is possible due to the nonlinearity of the self-dual equation. Due to the Bianchi identity, \begin{eqnarray} {\cal D}_{\nu} [{\cal A}]{\cal F}_{\mu\nu}^*[{\cal A}] \equiv 0 , \end{eqnarray} any instanton (anti-instanton) solution is also a solution of the Yang-Mills field equation, \begin{eqnarray} {\cal D}_{\nu} [{\cal A}]{\cal F}_{\mu\nu}[{\cal A}] = 0 , \end{eqnarray} but the converse does not hold. In fact, the instanton and the anti-instanton do not exhaust the solution of the Yang-Mills field equation, since there exists at least one solution of the Yang-Mills field equation (Sibner-Sibner-Uhlenbeck (SSU) solution \cite{SSU89}) which is not a solution of the self-dual equation, see Fig.~\ref{fig:moduli}. The existence of the instanton solution is suggested from the non-triviality of Homotopy group \cite{NS83,Nakahara90,Mermin79} $\pi_3(G)$, \begin{eqnarray} \pi_3(SU(N)) = {\bf Z} \ (N=2,3, \cdots) . \end{eqnarray} It is possible to construct the instanton background by choosing the gauge fixing condition, \begin{eqnarray} F^A[\Omega] = {\cal F}_{\mu\nu}^{\pm}{}^A[\Omega], \quad {\cal F}_{\mu\nu}^{\pm}{}^A[\Omega] := {\cal F}_{\mu\nu}^A[\Omega] \mp {\cal F}_{\mu\nu}^A{}^*[\Omega] . \label{SDYM} \end{eqnarray} It is shown \cite{KN98} that this choice leads to the topological Yang-Mills theory \cite{Witten88,TQFT} which is an example of the TQFT of Witten type. Since the topological Yang-Mills theory is derived from the N=2 Supersymmetric Yang-Mills theory by the procedure called the twisting, this might shed more light on the quark confinement based on the dual Meissner effect or the magnetic monopole \cite{SW94}. However, quark confinement will be realized only when the $N=2$ supersymmetry is broken down to $N=1$ by adding the mass perturbation. Since we do not have any convincing argument to justify such a scenario, we do not consider this possibility anymore in this paper. \subsection{Magnetic monopole current} \par In our formulation, however, the background field $\Omega_\mu$ is not a priori required to be the classical solution of the field equation (\ref{Feq}), when we consider the quantum theory of the background field $\Omega_\mu$. In quantum theory, it is not necessarily true that the most dominant contribution is given by the solution of the field equation. This is obvious in the functional integral approach because we must take into account the entropy associated with the relevant field configurations, which comes from the integration measure $[d\Omega_\mu]$ of the functional integral. In fact, whether the phase transition occurs or not is determined according to the balance between the action (energy) and the entropy, which is called the action (energy)-entropy argument. What kind of field configuration is important may vary from problem to problem. \par In our approach, we take the magnetic monopole current as the topologically nontrivial background $\Omega_\mu$. This choice is suggested from the recent result \cite{SY90} of Monte Carlo simulations in lattice gauge theories; the (Abelian) magnetic monopole after Abelian projection \cite{tHooft81} plays the dominant role in quark confinement. This fact is called the (Abelian) magnetic monopole dominance \cite{EI82}. \par The magnetic monopole in pure Yang-Mills theory (without the elementary Higgs scalar field) is obtained as follows. First, we restrict the Non-Abelian gauge group $G$ to the subgroup $H$ ($G \rightarrow H$) and retain only the gauge invariance for $H$, in other words, the gauge group element $U(x) \in G$ is restricted to the coset $G/H$. To obtain Abelian magnetic monopole, $H$ is chosen to be the maximal torus subgroup of $G$ (We will discuss other choices in the final section). We realize this restriction by the partial gauge fixing. The MAG is a partial gauge fixing so that $G/H$ is fixed and $H$ is retained by choosing $F^A[\Omega]$ appropriately. The MA gauge condition is obtained by minimizing the ${\cal R}[{\cal A}^U]$ with respect to the gauge rotation $U$ where \begin{eqnarray} {\cal R}[{\cal A}] := \int d^Dx \ {\rm tr}_{G/H} \left( {1 \over 2}{\cal A}_\mu(x) {\cal A}_\mu(x) \right) \equiv \int d^Dx \ {1 \over 2}A_\mu^a(x) A_\mu^a(x) , \end{eqnarray} where we have used the Cartan decomposition which decomposes the non-Abelian gauge field into the diagonal and the off-diagonal pieces, \begin{eqnarray} {\cal A}_\mu = {\cal A}_\mu^A T^A = a_\mu^\alpha T^\alpha + A_\mu^a T^a . \label{Cartandecomp} \end{eqnarray} Note that the trace is taken only on the coset part, see Appendix B. A geometric meaning of this function is given in section 7. According to the Cartan decomposition, the Abelian gauge potential is defined by \begin{eqnarray} a_\mu^\alpha(x) := {\rm tr}[{\cal H}_\alpha {\cal A}_\mu(x)] \end{eqnarray} where ${\cal H}_\alpha=T^\alpha (i=1, \cdots, {\rm rank}G)$ is the Cartan subalgebra. For $G=SU(2)$, the differential MA gauge is obtained as \begin{eqnarray} F^{a}[{\cal A}] := (\partial_\mu \delta^{ab} - \epsilon^{ab3} A_\mu{}^3) A_\mu^b := D_\mu{}^{ab}{}[A^3] A_\mu^b \quad (a,b = 1,2) , \label{dMAG} \end{eqnarray} where $a_\mu=A_\mu^3$. Note that the equation \begin{eqnarray} D_\mu{}^{ab}{}[A^3] A_\mu^b = 0 \quad (a,b = 1,2) . \label{MAg0} \end{eqnarray} is a 1st order nonlinear partial differential equation. We call this equation the monopole equation in what follows. \par Next, using the solution of $F^{a}[{\cal A}]=0$, the magnetic monopole current is defined by \begin{eqnarray} k_\mu^\alpha := \partial_\nu \tilde f_{\mu\nu}^\alpha = {1 \over 2} \epsilon_{\mu\nu\rho\sigma} \partial_\nu f_{\rho\sigma}^\alpha, \end{eqnarray} where the Abelian field strength is given by \begin{eqnarray} f_{\mu\nu}^\alpha := \partial_\mu a_\nu^\alpha - \partial_\nu a_\mu^\alpha . \end{eqnarray} Due to the topological conservation law, $ \partial_\mu k_\mu^\alpha = 0 $ the magnetic monopole current denotes a closed loop in four dimensions. The respective magnetic monopole is characterized by an integer-valued topological (magnetic) charge $Q_m^\alpha := \int d^3x k_0^\alpha(x) \in {\bf Z}$ ($\alpha=1,\cdots,N-1={\rm rank}SU(N)$). For the static monopole, the monopole current is given by $k_\mu^\alpha(x)=Q_m^\alpha \delta^3({\bf x}) \delta_{\mu 0}$ with $Q_m^\alpha$ being the magnetic charge. \par Finally, we must check the finiteness of ${\cal R}[{\cal A}]$, which is necessary to define the Morse function, see section 7. The instanton solution give a finite Yang-Mills action, i.e., $S_{YM}[{\cal A}]<\infty$ irrespective of the gauge choice, as a consequence of self-duality of the equation. On the other hand, the magnetic monopole solution of $F^{a}[{\cal A}]=0$ must give a finite ${\cal R}[{\cal A}]$, i.e., ${\cal R}[{\cal A}]<\infty$. This condition leads to the finiteness of the gauge-fixing action, \begin{eqnarray} S_{GF}[\Omega] < \infty , \end{eqnarray} in the MA gauge. In the gauge-fixed formulation of the quantum gauge field theory, the gauge fixing part $S_{GF}$ is important as well as the Yang-Mills action, $S_{YM}$. In what folllows, it is very convenient to separate the pure gauge piece in the gauge potential, \begin{eqnarray} \tilde \Omega_\mu(x) := {i \over g} \tilde U(x) \partial_\mu \tilde U^\dagger(x) = \tilde \Omega_\mu^A(x) T^A , \quad \tilde U \in G/H . \end{eqnarray} \par For $G=SU(2)$ and $H=U(1)$, it is shown that the Abelian gauge potential calculated as \begin{eqnarray} a_i(x) = {\rm tr}[{1 \over 2}\sigma_3 \tilde \Omega_i(x)] (i=1,2,3) \end{eqnarray} agrees exactly with the well-known static potential for the Dirac magnetic monopole \cite{Dirac31,WY75,GO78}, see e.g. \cite{KondoI}. \par From the mathematical point of view, the existence of magnetic monopole is consistent with the following relation for the Homotopy group, \begin{eqnarray} \pi_2(G/H) = \pi_1(H) \ {\rm when} \ \pi_2(G)=0 . \end{eqnarray} Usually, the pure Yang-Mills theory does not have magnetic monopole as a stable topological soliton. This is consistent with \begin{eqnarray} \pi_2(G) = 0 . \end{eqnarray} Therefore, for the existence of the magnetic monopole in pure Yang-Mills theory, the coset structure $G/H$ is an indispensable ingredient. For $G=SU(N)$, magnetic monopoles of $N-1$ species are expected for the maximal torus group $H=U(1)^{N-1}$, since \begin{eqnarray} \pi_2(SU(N)/U(1)^{N-1}) = \pi_1(U(1)^{N-1}) = {\bf Z}^{N-1} , \end{eqnarray} whereas \begin{eqnarray} \pi_2(SU(N)) = 0 \ (N=2,3, \cdots) . \end{eqnarray} \par \section{Deformation of a topological field theory} \setcounter{equation}{0} \subsection{Change of field variables} \par For the decomposition of field variable, \begin{eqnarray} {\cal A}_\mu(x) = \Omega_\mu(x) + {\cal Q}_\mu(x) , \label{deco} \end{eqnarray} it is possible to identify the gauge transformation \begin{eqnarray} \delta {\cal A}_\mu(x) = {\cal D}_\mu[A] {\omega}(x) := \partial_\mu {\omega}(x) - ig [{\cal A}_\mu(x), {\omega}(x)], \end{eqnarray} with a set of transformations \begin{eqnarray} \delta \Omega_\mu(x) &=& {\cal D}_\mu[\Omega] \omega(x) , \label{traOme3} \\ \delta {\cal Q}_\mu(x) &=& i g [\omega(x), {\cal Q}_\mu(x)] . \label{traQ3} \end{eqnarray} Here (\ref{traOme3}) and (\ref{traQ3}) correspond to (\ref{traOme}) and (\ref{traQ}) respectively. Note that $\Omega_\mu$ transforms as an gauge field, while ${\cal Q}_\mu$ as a adjoint matter field. \par When ${\cal A}_\mu$ is given by a finite gauge rotation (large gauge transformation) $U(x)$ of ${\cal V}_\mu$, we take the following identification, \begin{eqnarray} \Omega_\mu(x) := {i \over g} U(x) \partial_\mu U^\dagger(x), \quad {\cal Q}_\mu(x) := U(x) {\cal V}_\mu(x) U^\dagger(x) , \label{cov} \end{eqnarray} where we have identified $\Omega_\mu(x)$ with the background field which is supposed to be generated from $U(x)$. This identification leads after simple calculation to \begin{eqnarray} {\cal D}_\mu[\Omega] {\cal Q}_\mu &:=& \partial_\mu {\cal Q}_\mu - i g[\Omega_\mu, {\cal Q}_\mu] \\ &=& \partial_\mu (U(x) {\cal V}_\mu(x) U^\dagger(x)) + [ U(x) \partial_\mu U^\dagger(x), U(x) {\cal V}_\mu(x) U^\dagger(x) ] \nonumber\\ &=& U(x) \partial_\mu {\cal V}_\mu(x) U^\dagger(x) . \label{gft} \end{eqnarray} Therefore, the BGF gauge for ${\cal Q}_\mu$, \begin{eqnarray} {\cal D}_\mu[\Omega] {\cal Q}_\mu(x) = 0 , \label{BGFG} \end{eqnarray} is equivalent to the Lorentz gauge for ${\cal V}_\mu$, \begin{eqnarray} \partial_\mu {\cal V}_\mu(x) = 0 , \label{Lorentz} \end{eqnarray} under the identification of the variables (\ref{cov}). Under (\ref{cov}), we can rewrite (\ref{traOme3}) as \begin{eqnarray} \delta U \partial_\mu U^\dagger + U \partial_\mu \delta U^\dagger = ig \omega U \partial_\mu U^\dagger - ig U \partial_\mu(U^\dagger \omega) , \end{eqnarray} and (\ref{traQ3}) as \begin{eqnarray} \delta U {\cal V}_\mu U^\dagger + U {\cal V}_\mu \delta U^\dagger + U \delta {\cal V}_\mu U^\dagger = ig \omega U {\cal V}_\mu U^\dagger - ig U {\cal V}_\mu U^\dagger \omega . \end{eqnarray} Therefore, in the BGF gauge, (\ref{traOme3}) and (\ref{traQ3}) reduce to a set of transformations, \begin{eqnarray} \delta U(x) &=& i g \omega(x) U(x) , \quad \delta U^\dagger(x) = - i g U^\dagger(x) \omega(x) , \label{traU} \\ \delta {\cal V}_\mu(x) &=& 0 , \label{traV} \end{eqnarray} since the gauge degrees of freedom for ${\cal V}_\mu$ (small gauge transformation) is fixed by the Lorentz gauge (\ref{Lorentz}). In what follows, we assume that the non-compact gauge field variable ${\cal V}_\mu(x)$ does not have topologically nontrivial configuration and all topologically nontrivial contributions come from the compact gauge group variable $U(x)$ alone. We treat ${\cal V}_\mu(x)$ and $U(x)$ as if they are independent variables. The topological soliton (magnetic monopole) is derived as a solution of the nonlinear equation for $\Omega_\mu$ which follows from the nonlinear gauge fixing condition (MA gauge). The local gauge invariance of $\tilde Z[J,\Omega]$ written in terms of $\Omega_\mu$ and ${\cal Q}_\mu$ reduces to the invariance under the transformation (\ref{traU}), i.e. \begin{eqnarray} U(x) \rightarrow e^{ig\omega(x)}U(x) . \label{grot} \end{eqnarray} The measure $[d\Omega]$ invariant under (\ref{traOme}) is replaced by the invariant Haar measure $[dU]$ which is invariant under the local gauge rotation (\ref{grot}). \subsection{BRST formalism} First, we rewrite the BRST formulation of BGFM in terms of new variables. By making the change of variable (\ref{cov}) which is a gauge transformation of ${\cal V}(x)$ by $U(x)$, it turns out that the BRST transformation (\ref{BRST0}) for the variables $\Omega_\mu, {\cal Q}_\mu, \bar {\tilde C}, \bar {\tilde C}, \tilde B$ is rewritten into \begin{eqnarray} \tilde \delta_B U(x) &=& 0 , \nonumber\\ \tilde \delta_B {\cal V}_\mu(x) &=& {\cal D}_\mu[{\cal V}] \gamma(x) , \nonumber\\ \tilde \delta_B \gamma(x) &=& i{1 \over 2}g[\gamma(x), \gamma(x)], \nonumber\\ \tilde \delta_B \bar \gamma(x) &=& i \beta(x) , \nonumber\\ \tilde \delta_B \beta(x) &=& 0 , \label{BRST1} \end{eqnarray} where ${\cal V}_\mu, \gamma, \bar \gamma, \beta$ are the adjoint rotation of ${\cal Q}_\mu, \tilde C, \bar {\tilde C}, \bar B$ respectively, \begin{eqnarray} {\cal V}_\mu := U^\dagger {\cal Q}_\mu U, \quad \gamma := U^\dagger \tilde C U, \quad \bar \gamma := U^\dagger \bar {\tilde C} U, \quad \beta := U^\dagger \tilde B U . \label{adjr} \end{eqnarray} Under the adjoint rotation (\ref{adjr}), the measure is invariant, \begin{eqnarray} [d{\cal V}] [d\gamma][d\bar \gamma][d\beta] = [d{\cal Q}] [d\tilde C][d\bar {\tilde C}][d\tilde B] . \end{eqnarray} The Yang-Mills action is invariant \footnote{ For the definition of the field strength \begin{eqnarray} {\cal F}_{\mu\nu}[{\cal A}] := \partial_\mu {\cal A}_\nu - \partial_\nu {\cal A}_\mu - i g [ {\cal A}_\mu, {\cal A}_\nu] , \end{eqnarray} the change of variable ${\cal A}_\mu = U{\cal V}_\mu U^\dagger + { i \over g} U \partial_\mu U^\dagger$ leads to \begin{eqnarray} {\cal F}_{\mu\nu}[{\cal A}] = U {\cal F}_{\mu\nu}[{\cal V}] U^\dagger + { i \over g} U [\partial_\mu, \partial_\nu] U^\dagger . \end{eqnarray} Note that the second term ${\cal F}_{\mu\nu}^s :={ i \over g}U[\partial_\mu, \partial_\nu] U^\dagger$ can have a nonzero value and may modify the action. If $U(x) \in G$ is restricted to the coset $G/H$, it may yield a line-like singularity. For example, it is possible to have ${\cal F}_{xy}^s := {2\pi \over g} \delta(x)\delta(y)\theta(z) \sigma_3$ which corresponds to the presence of Dirac string extending into the direction of the negative $z$ axis from the origin, see e.g. Appendix C of \cite{KondoI}. Here the factor ${2\pi \over g}$ corresponds to the magnetic charge. The same contribution as the Dirac string can be incorporated by taking into account the magnetic monopole instead of the Dirac string, as shown in \cite{KondoI}. Moreover, the existence of such terms introduces rather singular terms in the action. Thus, we do not consider the effect of this term in what follows. } under this change of variables, \begin{eqnarray} S_{YM}[{\cal A}] = S_{YM}[\Omega+{\cal Q}] = S_{YM}[{\cal V}] . \end{eqnarray} The gauge fixing part (\ref{GF0}) for the BGF gauge is transformed into \begin{eqnarray} \tilde S_{GF}[{\cal V}, \gamma, \bar \gamma, \beta] &:=& - \int d^Dx \ i \tilde \delta_B \ {\rm tr}_G \left[ \bar \gamma \left( \partial_\mu {\cal V}_\mu +{\tilde \alpha \over 2}\beta \right) \right] \\ &=& \int d^Dx \ {\rm tr}_G \left[ \beta \partial_\mu {\cal V}_\mu + {\tilde \alpha \over 2} \beta \beta + i {\bar \gamma} \partial_\mu {\cal D}_\mu[{\cal V}] \gamma \right] , \label{GF2} \end{eqnarray} where we have used (\ref{gft}) and (\ref{adjr}). It turns out that the gauge fixing condition for ${\cal V}_\mu$ field is given by the Lorentz gauge (\ref{Lorentz}). Note that (\ref{GF2}) agrees with the form given in \cite{KondoII}. This BRST transformation corresponds to the small gauge transformation which does not change the topology of the gauge field. Thus the generating functional (\ref{ZBGFM}) is transformed as \begin{eqnarray} \tilde Z[J, \Omega] = \int [d{\cal V}] [d\gamma][d\bar \gamma][d\beta] \exp \left\{ i [ S_{YM}[{\cal V}] + \tilde S_{GF}[{\cal V}, \gamma, \bar \gamma, \beta] + (J_\mu \cdot U {\cal V}_\mu U^\dagger) ] \right\} \end{eqnarray} \par Next, we consider the total generating functional \begin{eqnarray} && Z[J] \nonumber\\&=& \int [d\Omega_\mu][dC][d\bar C][dB] \tilde Z[J, \Omega] \exp (i S_{GF}[\Omega, C, \bar C, B]) \exp [i(J_\mu \cdot \Omega_\mu)] \\ &=& \int [d\Omega_\mu] [dC][d\bar C][dB] \exp \{ i \tilde S_{eff}[J, \Omega] + i S_{GF}[\Omega, C, \bar C, B] +i(J_\mu \cdot \Omega_\mu) \} . \end{eqnarray} where we have introduced the source term $(J_\mu \cdot \Omega_\mu)$ for the background field. We introduce the BRST transformation, \begin{eqnarray} \delta_B \Omega_\mu(x) &=& {\cal D}_\mu[\Omega] C(x) := \partial_\mu C(x) - i g[\Omega_\mu(x), C(x)], \nonumber\\ \delta_B C(x) &=& i{1 \over 2}g[C(x), C(x)], \nonumber\\ \delta_B \bar C(x) &=& i B(x) , \nonumber\\ \delta_B B(x) &=& 0 , \label{BRST2} \end{eqnarray} and the anti-BRST transformation, \begin{eqnarray} \bar \delta_B \Omega_\mu(x) &=& {\cal D}_\mu[\Omega] \bar C(x) := \partial_\mu \bar C(x) - i g[\Omega_\mu(x), \bar C(x)], \nonumber\\ \bar \delta_B \bar C(x) &=& i{1 \over 2}g[\bar C(x), \bar C(x)], \nonumber\\ \bar \delta_B C(x) &=& i \bar B(x) , \nonumber\\ \bar \delta_B \bar B(x) &=& 0 , \label{BRST3} \end{eqnarray} where $\bar B$ is defined by \begin{eqnarray} B(x) + \bar B(x) = g [C(x), \bar C(x)] . \end{eqnarray} The BRST and anti-BRST transformations have the following properties, \begin{eqnarray} (\delta_B)^2 = 0, \quad (\bar \delta_B)^2 = 0, \quad \{ \delta_B, \bar \delta_B \} := \delta_B \bar \delta_B + \bar \delta_B \delta_B = 0 . \label{nilpotent} \end{eqnarray} \par In what follows, we consider the variable $U(x)$ as the fundamental variable instead of $\Omega_\mu(x)$. Then the BRST and anti-BRST transformations for $U$ and ${\cal V}_\mu$ are given by \begin{eqnarray} \delta_B U(x) = i g C(x) U(x), \quad \bar \delta_B U(x) = i g \bar C(x) U(x) , \label{UBRST} \end{eqnarray} and \begin{eqnarray} \delta_B {\cal V}_\mu(x) = 0 = \bar \delta_B {\cal V}_\mu(x) , \label{VBRST} \end{eqnarray} which are the BRST version of (\ref{traU}) and (\ref{traV}) respectively. In fact, (\ref{UBRST}) reproduces the usual BRST (\ref{BRST2}) and anti-BRST (\ref{BRST3}) transformations of the gauge field, \begin{eqnarray} \Omega_\mu(x) := {i \over g} U(x) \partial_\mu U^\dagger(x) . \end{eqnarray} Note that (\ref{UBRST}) and (\ref{VBRST}) lead to \begin{eqnarray} \delta_B \Omega_\mu(x) &=& {\cal D}_\mu[\Omega] C(x) , \\ \delta_B {\cal Q}_\mu(x) &=& i g [C(x), {\cal Q}_\mu(x)] . \end{eqnarray} These are the BRST version of (\ref{traOme3}) and (\ref{traQ3}) respectively, since within the BGFM, $\Omega_\mu, C, \bar C, B$ or $U, C, \bar C, B$ are external fields in the sense that they are not integrated out in the measure $[d{\cal V}] [d\tilde C][d\bar {\tilde C}][d\tilde B]$. Thus the generating functional of the total Yang-Mills theory reads \begin{eqnarray} Z[J] = \int [dU] [dC][d\bar C][dB] \exp \{ i \tilde S_{eff}[J, U] + i S_{GF}[\Omega, C, \bar C, B] +i(J_\mu \cdot \Omega_\mu) \} , \end{eqnarray} where $[dU]$ is the invariant Haar measure and we have redefined \begin{eqnarray} \tilde S_{eff}[J, U] := - i \ln \tilde Z[J, \Omega] . \end{eqnarray} \par In order to realize the magnetic monopole background, we adopt the MA gauge for which the gauge fixing and the FP ghost terms are written in the form \cite{KondoI} \begin{eqnarray} S_{GF}[\Omega, C, \bar C, B] &:=& - \int d^Dx \ i \delta_B \ {\rm tr}_{G/H} \left[ \bar C \left( F[\Omega]+{\alpha \over 2} B \right) \right] , \label{GF1} \end{eqnarray} where the trace is taken on the coset $G/H$, not the entire $G$. For $G=SU(2)$, \begin{eqnarray} F^{a}[\Omega] := (\partial^\mu \delta^{ab} - \epsilon^{ab3} \Omega^\mu{}^3) \Omega_\mu^b := D^\mu{}^{ab}{}[\Omega^3] \Omega_\mu^b \quad (a,b = 1,2) . \label{dMAG2} \end{eqnarray} \par By adding an BRST-exact ghost self-interaction term, (\ref{GF1}) is cast into the more convenient form \cite{KondoII}, \begin{eqnarray} S_{GF}'[\Omega, C, \bar C, B] := \int d^Dx \ i \delta_B \bar \delta_B {\rm tr}_{G/H} \left[ {1 \over 2} \Omega_\mu(x) \Omega_\mu(x) - {\alpha \over 2} i C(x) \bar C(x) \right] . \label{GF'} \end{eqnarray} From (\ref{nilpotent}), $S_{GF}'$ is invariant under the BRST and anti-BRST transformations, \begin{eqnarray} \delta_B S_{GF}' = 0 = \bar \delta_B S_{GF}' . \end{eqnarray} This action $S_{GF}'$ describes the topological soliton derived from the nonlinear equation $F[\Omega]=0$, the monopole equation. This action is BRST exact and hence there is no local degrees of freedom propagating in spacetime. It describes the quantity related to the global topology, as though the Chern-Simons theory describes the linking of knots \cite{Witten89}. We call the theory with the BRST exact action $S_{GF}'=S_{TQFT}$ alone the topological quantum field theory (TQFT). The generating functional is given by \begin{eqnarray} Z_{TQFT}[J] = \int [dU] [dC][d\bar C][dB] \exp \{ i S_{TQFT}[\Omega, C, \bar C, B] +i(J_\mu \cdot \Omega_\mu) \} . \label{TQFT} \end{eqnarray} In view of this, the above reformulation of the Yang-Mills theory was called the deformation of the TQFT. \footnote{ Different reformulations based on the similar idea have been presented by many authors, e.g., by Hata and Taniguchi \cite{HT95}, and Fucito, Martellini and Zeni \cite{FMZ97}. } \par In the above rederivation, the fact that the field $\Omega_\mu$ behaves as if it is a gauge field ${\cal A}_\mu$ is essential. This is guaranteed by the BGFM. \subsection{Expectation value} \par In our formulation, an arbitrary function $f({\cal A})$ of ${\cal A}$ is written as $f({\cal A}) = g({\cal V}_\mu, U) h(U)$ by making the change of variable (\ref{deco}) and (\ref{cov}). Then the expectation value is evaluated as \begin{eqnarray} \langle f({\cal A}) \rangle_{YM} = \langle \langle g({\cal V}_\mu, U) h(U) \rangle_{pYM}^{{\cal V}} \rangle_{TQFT}^U = \langle \langle g({\cal V}_\mu, U) \rangle_{pYM}^{{\cal V}} h(U) \rangle_{TQFT}^U . \label{expec} \end{eqnarray} Here, taking the expectation value $\langle \cdot \rangle_{TQFT}^U$ corresponds to summing over topological soliton contributions by making use of the TQFT described by the variable $U$, whereas $\langle \cdot \rangle_{pYM}^{{\cal V}}$ denotes the expectation value for the deformation piece which is described by the usual Yang-Mills theory with the variable ${\cal V}_\mu$ (Here $p$ denotes the perturbative). Of course, we can change the ordering of taking the expectation value, \begin{eqnarray} \langle f({\cal A}) \rangle_{YM} = \langle \langle g({\cal V}_\mu,U) h(U) \rangle_{TQFT}^{U} \rangle_{pYM}^{{\cal V}} . \end{eqnarray} Both expressions should give the same result, if they are calculated exactly. \par Under the assumption of perturbative deformation, the expectation $\langle g({\cal V}_\mu,U) \rangle_{pYM}$ is calculated by expanding the integrand $g({\cal V}_\mu,U)$ into power series in ${\cal V}_\mu$. This is a minimal assumption in the practical calculation. After that, $\langle g({\cal V}_\mu,U) \rangle_{pYM}$ is still a function of $U$, say, $p(U)$. Finally, the expectation $\langle p(U)h(U) \rangle_{TQFT}$ must be evaluated in the non-perturbative way, since this piece estimates the soliton contribution. Perturbative deformation is an assumption that the deformation part is evaluated in the perturbation theory in the coupling constant $g$. In other words, all the essential non-perturbative contributions are provided with the topological soliton described by the TQFT. Actually, this strategy was performed in the evaluation of the Wilson loop \cite{KondoII,KondoIV}. \subsection{Abelian-projected effective gauge theory} The above result should be compared with the previous formulation \cite{QR97,KondoI} which begins with the generating functional, \begin{eqnarray} Z[J] = \int [d{\cal A}_\mu] [dC][d\bar C][dB] \exp \{ i S_{YM}[{\cal A}] + i S_{GF}[{\cal A}, C, \bar C, B] +i(J_\mu \cdot {\cal A}_\mu) \} . \end{eqnarray} First, following the Cartan decomposition (\ref{Cartandecomp}), the non-Abelian gauge field was decomposed into the diagonal and the off-diagonal pieces, \begin{eqnarray} {\cal A}_\mu = {\cal A}_\mu^A T^A = a_\mu^i T^i + A_\mu^a T^a . \end{eqnarray} Then the MA gauge was imposed as a gauge fixing condition. Finally, all the off-diagonal fields taking values in the Lie algebra of the coset $G/H$ were integrated out in the functional integral, \begin{eqnarray} Z[J] = \int [da_\mu^i] [dC^i][d\bar C^i][dB^i] \exp \{ i S_{diag}[a^i,C^i,\bar C^i,B^i] +i(J_\mu \cdot a_\mu)\} , \end{eqnarray} where \begin{eqnarray} && Z[a^i,C^i,\bar C^i,B^i] := \exp \{ i S_{diag}[a^i,C^i,\bar C^i,B^i] \} \\ &&:= \int [dA_\mu^a] [dC^a][d\bar C^a][dB^a] \exp \{ i S_{YM}[{\cal A}] + i S_{GF}[{\cal A}, C, \bar C, B] +i(J_\mu \cdot A_\mu) \} \end{eqnarray} The theory with the action $S_{diag}[a^i,C^i,\bar C^i,B^i]$ was called the Abelian-projected effective gauge theory (APEGT). It has been shown \cite{QR97,KondoI} that the APEGT has the same beta function as the original Yang-Mills theory, exhibiting the asymptotic freedom, although the APEGT is an Abelian gauge theory. \par It turns out that the previous strategy presented in \cite{QR97,KondoI} is equivalent to the above formulation presented in this paper and that the results obtained in the previous works are the immediate consequence of the present formulation, if we identify the diagonal and off-diagonal fields with the background field and the quantum fluctuation respectively, i.e., \begin{eqnarray} \Omega_\mu = a_\mu^i T^i, \quad {\cal Q}_\mu = A_\mu^a T^a . \end{eqnarray} The theory with an action $S_{diag}[a^i,C^i,\bar C^i,B^i]$ is written in terms of only the diagonal fields. As long as the BGF gauge is imposed on the off-diagonal field $A_\mu$, this theory becomes the Abelian gauge theory, since the BGFM guarantees that the background field $a_\mu$ transforms as a gauge field (Of course, the diagonal field is reduced to the Abelian gauge field in this case). Indeed, the BGF gauge $D^{ab}[a]A^b=0$ is nothing but the MA gauge. Hence the coincidence of the beta function is understood from the BGFM. \section{Strategy of a derivation of quark confinement} \setcounter{equation}{0} We consider the D-dim. QCD (QCD$_D$) with a gauge group G for $D >2$. The (full) non-Abelian Wilson loop is defined as the path-ordered exponent along a loop $C$, \begin{eqnarray} W^C [{\cal A}] := {\rm tr} \left[ {\cal P} \exp \left( i g \oint_C {\cal A}_\mu^A(x) T^A dx^\mu \right) \right] /{\rm tr}(1) . \end{eqnarray} We define the (full) string tension $\sigma$ by \begin{eqnarray} \sigma := - \lim_{A(C) \rightarrow \infty} {1 \over A(C)} \ln \langle W^C[{\cal A}] \rangle , \end{eqnarray} where $A(C)$ is the minimal area spanned by the Wilson loop $C$. The non-zero string tension $\sigma \not=0$ implies that the Wilson loop expectation value behaves for large loop as \begin{eqnarray} \langle W^C[{\cal A}] \rangle \sim \exp (- \sigma A(C)) . \end{eqnarray} This is called the area (decay) law. The static potential $V(R)$ for a pair of quark and anti-quark is evaluated from the rectangular Wilson loop $C$ with sides $T$ and $R$ ($A(C)=TR$) according to \begin{eqnarray} V(R) = - \lim_{T \rightarrow \infty} {1 \over T} \ln \langle W^C[{\cal A}] \rangle . \label{st} \end{eqnarray} The area law of the Wilson loop or non-zero string tension $\sigma \not=0$ implies the existence of the linear part $\sigma R$ in the static potential $V(R)$, leading to quark confinement. \par In a series of papers \cite{KondoI,KondoII,KondoIII,KondoIV,KondoV}, a derivation of the area law of the Wilson loop in $QCD_4$ has been given in the following steps. \begin{enumerate} \item[] Step 1: Reformulating the Yang-Mills theory as a deformation of a TQFT in MA gauge \cite{KondoII} \item[] Step 2: Parisi-Sourlas Dimensional reduction \cite{KondoII} \item[] Step 3: Abelian magnetic monopole dominance \cite{KondoIV} \item[] Step 4: Instanton calculus \cite{KondoII} or large $N$ expansion \cite{KT99} \end{enumerate} The first two steps are shown schematically as follows. \par \vskip 0.5cm \begin{center} \unitlength=1.0cm \thicklines \begin{picture}(12,6) \put(2,5){\framebox(8,1){D-dim. QCD with a gauge group $G$}} \put(6.2,5){\vector(0,-1){0.8}} \put(7,4.5){MA gauge} \put(-0.2,2.8){\framebox(12.4,1.4){}} \put(0,3){\framebox(5,1){D-dim. Perturbative QCD}} \put(6,3.5){$\bigotimes$} \put(5.5,3){deform} \put(7,3){\framebox(5,1){D-dim. TQFT}} \put(8.5,3){\vector(0,-1){1}} \put(9,2.4){Dimensional reduction} \put(-0.2,0.8){\framebox(12.4,1.4){}} \put(0,1){\framebox(5,1){D-dim. Perturbative QCD}} \put(6,1.5){$\bigotimes$} \put(5.5,1){deform} \put(7,1){\framebox(5,1){(D-2)-dim. G/H NLSM}} \end{picture} \end{center} The following analyses within the above reformulation (the perturbative deformation of a TQFT) is based on an assumption that the contribution from ${\cal V}_\mu(x)$ can be treated in perturbation theory in the gauge coupling constant $g$. This assumption leads to the topological sector dominance in the sense that the area law contribution comes solely from the contribution of $U(x)$ described by the TQFT and the remaining part ${\cal V}_\mu(x)$ does not contribute to the area law of the Wilson loop. \subsection{Step 1: Reformulating the Yang-Mills theory as a deformation of a TQFT in MA gauge} QCD$_D$ is reformulated as a deformation of a TQFT$_D$ in MA gauge. The MA gauge is a partial gauge fixing such that the coset part $G/H$ of the gauge group $G$ is fixed and the maximal torus group $H$ is left as a residual gauge group. \par For $G=SU(2)$, it has been shown \cite{KondoIV} that the expectation value of the non-Abelian Wilson loop is rewritten using the non-Abelian Stokes theorem \cite{KondoIV} into \begin{eqnarray} && \langle W^C[{\cal A}] \rangle_{YM} \nonumber\\ &=& \Biggr\langle \Biggr\langle \exp \left[ i g J \oint_C dx^\mu n^A(x) {\cal V}_\mu^A(x) \right] \Biggr\rangle_{pYM} \exp \left[ iJ \int_{S} d^2z \ \epsilon_{\mu\nu} {\bf n} \cdot (\partial_\mu {\bf n} \times \partial_\nu {\bf n}) \right] \Biggr\rangle_{TQFT} , \end{eqnarray} where $S$ is a surface with a boundary $C$ ($\partial S=C$) and ${\bf n}(x)=(n^1(x),n^2(x),n^3(x))$ is the three-dimensional unit vector (${\bf n}(x) \cdot {\bf n}(x) = 1$) defined by \begin{eqnarray} n^A(x) T^A = U^\dagger(x) T^3 U(x) , \quad T^A= {1 \over 2}\sigma^A \ (A=1,2,3) . \label{aop} \end{eqnarray} Here $J$ specifies the representation of the fermion in the definition of the Wilson loop and $J=1/2$ corresponds to the fundamental representation. \subsection{Step 2: Parisi-Sourlas dimensional reduction} It has been shown \cite{KondoII} that TQFT$_D$ is equivalent to the coset $G/H$ nonlinear sigma model (NLSM) in (D-2) dimensions, NLSM$_{D-2}$. This is a consequence of Parisi-Sourlas dimensional reduction \cite{PS79} due to the supersymmetry hidden in TQFT (\ref{TQFT}). This is an advantage that we have chosen the MA gauge. \par As extensively discussed more than 20 years ago, QCD$_4$ and NLSM$_2$ have various common properties: renormalizability, asymptotic freedom (i.e., negative beta function $\beta(g) < 0$), dynamical mass generation, existence of instanton solution, no phase transition for any value of coupling constant (i.e., one phase), etc. This similarity between two theories can be understood from this correspondence, \begin{eqnarray} QCD_4 \supset TQFT_4 \Longleftrightarrow G/H~ NLSM_2 . \end{eqnarray} For the $SU(N)$ gauge group, the existence of 2D instanton is guaranteed for any $N$, because $\pi_2(SU(N)/U(1)^{N-1})=\pi_1(U(1)^{N-1})={\bf Z}^{N-1}$. See the second paper in \cite{KT99} for details. \par For $G=SU(2)$, G/H NLSM is nothing but the O(3) NLSM. For the {\it planar} Wilson loop, the evaluation of the expectation value $\langle \cdot \rangle_{TQFT}$ in TQFT$_4$ \begin{eqnarray} \langle W^C[{\cal A}] \rangle_{YM} = Z_{TQFT_4}^{-1}\int [dU(x)]_{x \in {\bf R}^4} \exp (-S_{TQFT_4}[U] ) p(U)h(U) \end{eqnarray} is reduced to that in the coset $G/H$ NLSM$_2$ \begin{eqnarray} \langle W^C[{\cal A}] \rangle_{YM} = Z_{NLSM_2}^{-1}\int [d{\bf n}(x)]_{x \in {\bf R}^2} \exp (-S_{NLSM_2}[{\bf n}] ) p(U)h(U) , \label{Wc} \end{eqnarray} where we have used the notation (\ref{expec}) with \begin{eqnarray} h(U) &:=& \exp \left[ iJ \int_{S} d^2z \ \epsilon_{\mu\nu} {\bf n} \cdot (\partial_\mu {\bf n} \times \partial_\nu {\bf n}) \right] , \\ p(U) &:=& \Biggr\langle \exp \left[ i g J \oint_C dx^\mu n^A(x) {\cal V}_\mu^A(x) \right] \Biggr\rangle_{pYM} . \end{eqnarray} In the original Lagrangian of QCD, the scalar field is not included as an elementary field, but it appears as a composite field according to (\ref{aop}). The unit vector ${\bf n}(x)$ plays the same role as the monopole scalar field $\phi(x)$ which describes the 't Hooft-Polyakov monopole \cite{tHooft74,Polyakov74}, for $G=SU(2)$, i.e., \begin{eqnarray} n^A(x) \leftrightarrow \hat \phi^A(x) := {\phi^A(x) \over |\phi(x)|} , \quad |\phi(x)| := \sqrt{\phi^A(x)\phi^A(x)} . \end{eqnarray} \subsection{Step 3: Abelian magnetic monopole dominance} The diagonal (or Abelian) string tension $\sigma_{Abel}$ is defined by \begin{eqnarray} \sigma_{Abel} := - \lim_{A(C) \rightarrow \infty} {1 \over A(C)} \ln \left\langle W^C[a^\Omega] \right\rangle_{TQFT_4} , \end{eqnarray} by making use of the diagonal Wilson loop, \begin{eqnarray} W^C[a^\Omega] &=& \exp \left( i g J \oint_C dx^\mu a_\mu^\Omega (x) \right), \\ \quad a_\mu^\Omega(x) &:=& \Omega_\mu^3(x) := {\rm tr}(T^3 \Omega_\mu(x)) , \quad \Omega_\mu(x) := {i \over g} U(x) \partial_\mu U(x)^\dagger . \label{dWl} \end{eqnarray} Owing to the dimensional reduction, we find \begin{eqnarray} \left\langle W^C[a^\Omega] \right\rangle_{TQFT_4} = \left\langle W^C[a^\Omega] \right\rangle_{NLSM_2} , \end{eqnarray} where \begin{eqnarray} \langle W^C[a^\Omega] \rangle_{NLSM_2} = Z_{NLSM_2}^{-1}\int [d{\bf n}(x)]_{x \in {\bf R}^2} \exp (-S_{NLSM_2}[{\bf n}] ) W^C[a^\Omega] . \label{Wc1} \end{eqnarray} Then it is shown that, in the limit of large Wilson loop, two string tensions agree with each other, $\sigma = \sigma_{Abel}$, since \begin{eqnarray} {1 \over A(C)} \left[ \ln \left\langle W^C[{\cal A}] \right\rangle_{YM_4} - \ln \left\langle W^C[a^\Omega] \right\rangle_{NLSM_2} \right] \downarrow 0 \quad (A(C) \uparrow \infty) , \end{eqnarray} if we identify the deformation with the perturbative one. Thus, for the large planar (non-intersecting) Wilson loop , the full string tension $\sigma$ is saturated by the diagonal string tension $\sigma_{Abel}$. This explains the Abelian dominance and magnetic monopole dominance. \par This result is derived as follows. In calculating (\ref{Wc}), if we put $p(U)\equiv 1$, then $\left\langle W^C[{\cal A}] \right\rangle_{YM_4}$ coincides with $\left\langle W^C[a^\Omega] \right\rangle_{NLSM_2}$, since it is shown \cite{KondoII} that $h(U) \equiv W^C[a^\Omega]$. In our framework called the perturbative deformation of TQFT, $p(U)$ is estimated by making use of the power-series expansion in the coupling constant $g$ (or in the 't Hooft coupling $\lambda:= g^2N$ in the framework of large $N$ expansion, see \cite{KT99}) as \begin{eqnarray} p(U) &=& 1 + \sum_{n=1}^{\infty} {(i g)^n \over n!} \Biggr\langle \left( J \oint_C dx^\mu n^A(x) {\cal V}_\mu^A(x) \right)^n \Biggr\rangle_{pYM} \nonumber\\ &=& 1 + \sum_{n=1}^{\infty} {(i gJ)^n \over n!} \oint_C dx_1^{\mu_1} \cdots \oint_C dx_n^{\mu_n} n^{A_1}(x_1) \cdots n^{A_n}(x_n) \nonumber\\ && \times \langle {\cal V}_{\mu_1}^{A_1}(x_1) \cdots {\cal V}_{\mu_n}^{A_n}(x_n) \rangle_{pYM} . \end{eqnarray} By calculating the expectation value $\langle {\cal V}_{\mu_1}^{A_1}(x_1) \cdots {\cal V}_{\mu_n}^{A_n}(x_n) \rangle_{pYM}$, we can express $p(U)$ in terms of the ${\bf n}(x)$ fields which are defined on the loop $C$ embedded in the two-dimensional space. However, it is shown that the additional contribution from $p(U)-1$ to the expectation value of the Wilson loop does not have the the area decay part. Incidentally, although the perturbative deformation part is insufficient to derive the area law, it leads to the running coupling constant which is governed by the renormalization group $\beta$ function of the original Ynag-Mills theory in consistent with the asymptotic freedom. \subsection{Step 4: Instanton calculus} \par The whole problem is reduced to calculating the diagonal Wilson loop in NLSM$_2$, \begin{eqnarray} \left\langle W^C[a^\Omega] \right\rangle_{NLSM_2} = \left\langle e^{i 2\pi J Q_S } \right\rangle_{NLSM_2} , \quad Q_S = {1 \over 8\pi } \int_S d^2z \ \epsilon_{\mu\nu} {\bf n} \cdot (\partial_\mu {\bf n} \times \partial_\nu {\bf n}) . \end{eqnarray} Note that the integrand of $Q_S$ is the instanton density in NLSM$_2$. Therefore, $Q_S$ counts the number of instantons minus that of anti-instantons inside the area $S(\subset {\bf R}^2)$ bounded by the Wilson loop $C$. This suggests that the quark confinement follows from the condensation of topological soliton, the magnetic monopole. \par In this step we have employed the naive instanton calculus to calculate the diagonal Wilson loop. In the dilute gas approximation the two-dimensional instanton contributions are summing up according to \begin{eqnarray} \sum_{n_+=0}^{\infty} \sum_{n_-=0}^{\infty} {1 \over n_+! n_-!} \int \prod_{i=1}^{n} d^2z_i \int \prod_{i=1}^{n} d\mu(\rho_i) \exp [-(n_++n_-)S_1(g) ] e^{i 2\pi J Q_S } , \end{eqnarray} where the action of NLSM$_2$ is replaced by $(n_++n_-)S_1(g)$ using the numbers of instanton and anti-instanton $n_+, n_-$ and the action for one instanton $S_1(g)= 4\pi^2/g^2$ in NLSM$_2$. Thus the (infinite dimensional) functional integral measure $[d{\bf n}(x)]_{x\in {\bf R}^2}$ has been replaced with the (finite dimensional) integration with respect to the collective coordinates, ${z_i}$ (position of the instanton) and ${\rho_i}$ (size of the instanton). Such reduction of degrees of freedom in the functional integration is a common feature in TQFT as shown in section 7. \par This leads to the area law of the diagonal Wilson loop and the non-zero diagonal string tension $\sigma_{Abel}$ for half odd integer $J$ or the fractional charge $q$. \subsection{Area law and quark confinement} \par In the framework of the deformation of a TQFT for the Yang-Mills theory, the non-zero string tension $\sigma$ in QCD$_4$ follows from the non-zero diagonal string tension $\sigma_{Abel}$ in NLSM$_2$. The problem of proving area law in QCD$_4$ is reduced to the corresponding problem in NLSM$_2$. \par All the above steps are exact except for the instanton calculus of the Wilson loop in NLSM$_2$. For sufficiently large and planar Wilson loop, it was shown \cite{KondoII,KondoIV} that the string tension is given by \begin{eqnarray} \sigma = 2B e^{-S_1} \left[ 1 - \cos \left( 2\pi J \right) \right] , \quad S_1= {2\pi^2 \alpha \over g^2} , \label{str} \end{eqnarray} where $B$ is a constant with the mass-squared dimension coming from the integration over the instanton size $\int d\mu(\rho)$ and $S_1$ is the action for one instanton in NLSM$_2$. \par The result (\ref{str}) shows that for half odd integers $J={1 \over 2}, {3 \over 2}, {5 \over 2}, \cdots$, the Wilson loop exhibits area law for sufficiently large Wilson loop $C$, whereas the area law and the linear potential disappears for integers $J=1,2,3, \cdots$. Therefore, the fundamental fermion $J={1 \over 2}$ is confined, while the adjoint fermion $J=1$ can not be confined. \par \par In the above formulation using the MA gauge, it is the {\it compact} residual Abelian group that plays the essential role in evaluating the gauge invariant quantity. This feature is very similar to the situation in the lattice gauge theory. In fact, the result (\ref{str}) is a consequence of the periodicity (or compactness) of the residual Abelian gauge group, i.e., maximal torus group $U(1)$ of $SU(2)$, in the variable $U$ after the MA gauge is chosen. On the other hand, the gauge degrees of freedom for the non-compact field ${\cal V}_\mu$ have been completely fixed by the gauge fixing condition of Lorentz type. The explicit expression (\ref{str}) depends on the approximation taken in the instanton calculus, the periodicity of the string tension (hence the absence of string tension for $J=1,2,\cdots$) does not depend on the approximation. \par It is the perturbative part that gives the running of the coupling constant $g$. The running is governed by the renormalization group beta function $\beta(g)$. The usual Yang-Mills$_4$ theory exhibits asymptotic freedom, e.g., for $G=SU(N_c)$ at one-loop level, \begin{eqnarray} \beta(g) := \mu {dg(\mu) \over d\mu} = - {b_0 \over 16\pi^2} g(\mu)^3 + \cdots , \quad b_0 = {11N_c \over 3} > 0 . \end{eqnarray} In our framework, the correct beta function is derived based on the BGFM, see \cite{QR97,KondoI}. For the static potential $V(R)$, the perturbative part gives a Coulomb potential contribution $\alpha(\mu)/R$ where $\alpha(\mu):=g^2(\mu)/4\pi$ runs according to the $\beta(g)$. \par Similar strategy can also be applied to QED$_4$ ($G=U(1)$) to prove the existence of strong coupling confinement phase \cite{KondoIII}. This follows from the existence of Berezinski-Kosterlitz-Thouless transition of the O(2) NLSM$_2$. The corresponding steps are shown as follows. \par \vskip 0.5cm \begin{center} \unitlength=1cm \thicklines \begin{picture}(12,6) \put(2,5){\framebox(8,1){D-dim. QED}} \put(6.2,5){\vector(0,-1){0.8}} \put(7,4.5){Covariant Lorentz gauge fixing} \put(-0.2,2.8){\framebox(12.4,1.4){}} \put(0,3){\framebox(5,1){D-dim. Perturbative QED}} \put(6,3.5){$\bigotimes$} \put(5.5,3){deform} \put(7,3){\framebox(5,1){D-dim. TQFT}} \put(8.5,3){\vector(0,-1){1}} \put(9,2.4){Dimensional reduction} \put(-0.2,0.8){\framebox(12.4,1.4){}} \put(0,1){\framebox(5,1){D-dim. Perturbative QED}} \put(6,1.5){$\bigotimes$} \put(5.5,1){deform} \put(7,1){\framebox(5,1){(D-2)-dim. O(2) NLSM}} \end{picture} \end{center} This result enables us to give another derivation of quark confinement in QCD based on the low-energy effective {\it Abelian} gauge theory \cite{KondoI}, see \cite{KondoV}. This viewpoint is more interesting in the sense that the confinement-deconfinement transition can be discussed within the same framework. \subsection{Remarks and unresolved issues} \par The dilute gas approximation can be improved. More systematic instanton calculations enable us to identify an instanton solution with the Coulomb gas of vortices \cite{BL79,FFS79,BL81,Silvestrov90,Polyakov87}. Consequently, the low-energy effective Abelian gauge theory belongs to the strong coupling phase where the quark confinement is realized, see \cite{KondoV}. \par The absence of intermediate Casimir scaling region (i.e., $\sigma=0$ for integer $J$) may be due to our simplified treatment of the instanton size. In order to obtain the result (\ref{str}) we have treated the instanton as if it is exactly a point-like object in the dilute gas approximation. The Casimir scaling will be explained by taking into account the size effect of the instanton, as performed for the center vortex by Greensite et al. \cite{Greensite}. \par Recent investigations show that the QCD vacuum is a dual super conductor caused by the condensation of magnetic monopole and that the low-energy effective gauge theory is given the dual Ginzburg-Landau theory. However, numerical simulations claim that the dual superconductor is near type I, rather than type II, see \cite{Bali98}. This result seems to contradicts with the analytical studies. \par It is desirable to extend the above analyses into more general gauge groups. The case of $G=SU(3)$ will be discussed in forthcoming paper in detail. \subsection{A proposal of numerical calculations} Some of the implications from the above strategy will be checked by direct numerical simulations on the lattice. Due to difficulties of defining supersymmetry on the lattice, it might be impossible to check directly the equivalence between the 4D TQFT and the 2D NLSM. Nevertheless, it is desirable to check the following statements: \begin{enumerate} \item Validity of perturbative deformation of TQFT: The expectation value of the diagonal Wilson loop in NLSM$_2$, \begin{eqnarray} \left\langle W^C[a^\Omega] \right\rangle_{NLSM_2} = \left\langle \exp \left[ i 2\pi J {1 \over 8\pi } \int_S d^2z \ \epsilon_{\mu\nu} {\bf n} \cdot (\partial_\mu {\bf n} \times \partial_\nu {\bf n}) \right] \right\rangle_{NLSM_2} , \end{eqnarray} behaves as that of the non-Abelian Wilson loop in Yang-Mills$_4$, \begin{eqnarray} \left\langle W^C[{\cal A}] \right\rangle_{YM_4} = \left\langle {\rm tr} \left[ {\cal P} \exp \left( i g \oint_C {\cal A}_\mu^A(x) T^A dx^\mu \right) \right] \right\rangle_{YM_4} . \end{eqnarray} Two string tensions $\sigma_{Abel}$ and $\sigma$ agree with each other. \item Validity of instanton calculus: Only the instanton contribution in NLSM$_2$ is sufficient to recover the Abelian string tension, $\sigma_{Abel}$. \item Existence of the scale: The asymptotic scaling holds for the Abelian string tension $\sigma_{Abel}$ calculated from the NLSM$_2$. \end{enumerate} The results will prove or disprove validity of our strategy of deriving quark confinement. \section{Gauge fixing and gluon mass} \setcounter{equation}{0} \subsection{A naive MA gauge} A simple but ad hoc way to give the mass for the off-diagonal gluon is to introduce the following mass term to the Yang-Mills action, \begin{eqnarray} S_m = \int d^Dx {\rm tr}_{G/H} \left( {1 \over 2}m^2 {\cal A}_\mu {\cal A}_\mu \right) . \label{massterm} \end{eqnarray} This introduce the mass of the off-diagonal gluons in the tree level and this explicitly break the gauge invariance corresponding to $G/H$. Indeed, the mass term (\ref{massterm}) is derived as a gauge fixing term as follows. The simplest MA gauge where the off-diagonal part is made as small as possible will be the following gauge, \begin{eqnarray} F^a[{\cal A}] := A_\mu^a = 0 . \label{Agauge} \end{eqnarray} In order to write the gauge-fixing action, we must introduce the vector auxiliary field $B_\mu$ and the vector FP ghost $C_\mu$ and anti-ghost $\bar C_\mu$, so that \begin{eqnarray} S_{GF} = - \int d^Dx \ i\delta_B {\rm tr}_{G/H} \left[ \bar C_\mu \left( {\cal A}_\mu + {\alpha \over 2} B_\mu \right) \right] , \end{eqnarray} where the nilpotent BRST transformation is constructed as \footnote{ This BRST transformation does not leave the Yang-Mills action invariant, unless the equation of motion is used. So, it is unusual when we include the Yang-Mills action. } \begin{eqnarray} \delta_B {\cal A}_\mu &=& C_\mu, \nonumber\\ \delta_B C_\mu &=& 0, \nonumber\\ \delta_B \bar C_\mu &=& i B_\mu, \nonumber\\ \delta_B B_\mu &=& 0 . \end{eqnarray} Eliminating the auxiliary field $B_\mu$, we reproduce the mass term, \begin{eqnarray} S_{GF} = \int d^4x \ {\rm tr}_{G/H} \left[ {m^2_\alpha \over 2} {\cal A}_\mu(x){\cal A}_\mu (x) + i \bar C_\mu(x) C_\mu(x) \right] , \end{eqnarray} where we have put $m^2_\alpha=1/\alpha$. \footnote{ For $D=4$, the mass dimension is given as follows, dim[${\cal A}_\mu$] = dim[$C_\mu$] = 1, dim[$B_\mu$] = dim[$\bar C_\mu$] = 3 and dim[$\alpha$]=-2. } In four-dimensions, the parameter $m_\alpha$ looks like a mass which is arbitrary and can not be determined. The BRST transformation is highly unusual, since it corresponds to the gauge transformation much larger than the SU(N) gauge transformation. Note that $C_\mu^A$ has the same number of indices as ${\cal A}_\mu^A$. Hence we can use $C_\mu^A$ to eliminate the fields ${\cal A}_\mu^A$ to obtain the vacuous theory. The ghost field $C_\mu^A$ has its own remaining ghost symmetry, parameterized by the ghost field, $\phi^A$, the ghost for ghost, so the ghosts themselves require more gauge fixing. Note that the gauge fixing condition (\ref{Agauge}) does not allow the topological soliton, since it is linear in the field. \par Making the change of variables with the adjoint orbit parameterization \begin{eqnarray} n^A(x) = {\rm tr} \left[ U^\dagger(x) T^3 U(x)T^A \right] , \quad T^A= {1 \over 2}\sigma^A \ (A=1,2,3) \label{aop2} \end{eqnarray} lead to the four-dimensional coset G/H NLSM. \begin{eqnarray} S_{GF} = \int d^4x \ \left[ {m^2_\alpha \over 2} \partial_\mu {\bf n}(x)\partial_\mu {\bf n}(x) + \cdots \right] , \end{eqnarray} since ${\cal A}_\mu= U\partial_\mu U^\dagger + \cdots$. This is similar to a piece of the effective theory for the low-energy QCD proposed by Faddeev and Niemi \cite{FN98} based on Cho's works \cite{Cho80}. \par \subsection{The MA gauge} The naive MA gauge above should be compared with the MA gauge. The MA gauge \begin{eqnarray} F^{a}[\Omega] := (\partial^\mu \delta^{ab} - \epsilon^{ab3} \Omega^\mu{}^3) \Omega_\mu^b := D^\mu{}^{ab}{}[\Omega^3] \Omega_\mu^b \quad (a,b = 1,2) \label{dMAG3} \end{eqnarray} is obtained by minimizing the ${\cal R}[{\cal A}^U]$ with respect to the gauge rotation $U$ where \begin{eqnarray} {\cal R}[{\cal A}] := \int d^Dx \ {\rm tr}_{G/H} \left( {k \over 2}{\cal A}_\mu(x) {\cal A}_\mu(x) \right) , \end{eqnarray} where $k$ is a constant. The MA gauge fixing leads to the gauge-fixing action (\ref{GF1}) where the gauge fixing parameter is arbitrary at this stage. In our formulation, we demand the supersymmetry \cite{KondoII} of the gauge fixing action. Then the dimensional reduction \cite{PS79} occurs as a spontaneous breaking of the supersymmetry (as explained below). This symmetry requirement has determined the form of the gauge-fixing term (\ref{GF'}) and the result is independent from the coefficient $k$ in ${\cal R}[{\ cal A}]$. The explicit action (\ref{GF'}) after taking the BRST transformation is rather complicated and does not have any apparent mass term, see \cite{KondoI,KondoII}. However, the dimensional reduction \cite{KondoII} leads to \begin{eqnarray} S_{GF}' &:=& \int d^{D-2}z \ 2\pi {\rm tr}_{G/H} \left[ {1 \over 2} {\cal A}_a(z,{\bf 0}) {\cal A}_a(z,{\bf 0}) + i C(z,{\bf 0}) \bar C(z,{\bf 0}) \right] \\ &=& \int d^{D}x \ {\rm tr}_{G/H} \left[ {2\pi \over 2} {\cal A}_a(x) {\cal A}_a(x) + i 2\pi C(x) \bar C(x) \right] \delta^2(\hat x) , \label{GF''} \end{eqnarray} where $x=(z,\hat x) \in {\bf R}^D$ and $z \in {\bf R}^{D-2}$, $\hat x \in {\bf R}^2$, $a=1, \cdots, D-2$. Hence, the MA gauge leads to the unusual mass term, $m(x)=m(z,\hat x)=2\pi \delta^2(\hat x)$. The mass is anisotropic and the gauge field is massive only in $D-2$ dimensions. However, the choice of the (D-2)-dimensional subspace is arbitrary. For $D=4$, the equivalent action is given in the form of two-dimensional NLSM, \begin{eqnarray} S_{GF} = \int d^{2}z \ \left[ {2\pi/g^2 \over 2} \partial_a {\bf n}(z) \cdot \partial_a {\bf n}(z) + {\rm tr}_{G/H} \left( i 2 \pi \bar C_\mu(z) C_\mu(z) \right) \right] . \end{eqnarray} It is known that the two-dimensional NLSM exhibits dynamical mass generation, that is to say, the spectrum has a mass gap, although the initial lagrangian does not have the usual mass term. In this sense, in the subspace $R^2$ the gauge field can have the mass. \footnote{ The mass generation due to dimensional reduction to the NLSM was first demonstrated by Hata and Kugo \cite{HK85} in the context of color confinement. } \begin{figure} \begin{center} \unitlength=1cm \begin{picture}(13,6) \put(3.5,5.5){\bf $D=4+\epsilon$} \put(3.5,5.0){\bf $~d=2+\epsilon$} \put(0,5.5){\bf $\beta(g)$} \put(3,2.4){\bf $g_c \sim O(\epsilon)$} \thicklines \put(0,0){\vector(0,1){5}} \put(0,2){\vector(1,0){5}} \put(-0.5,1.9){\bf 0} \put(5.5,1.9){\bf $g$} \bezier{300}(0,2)(2,4)(5,0) \put(3,-0.5){\bf (a)} \put(10.5,5.5){\bf $D=4$} \put(10.5,5.0){\bf $~d=2$} \put(7,5.5){\bf $\beta(g)$} \thicklines \put(7,0){\vector(0,1){5}} \put(7,2){\vector(1,0){5}} \put(6.5,1.9){\bf 0} \put(12.5,1.9){\bf $g$} \bezier{300}(7,2)(9,2)(12,0) \put(10,-0.5){\bf (b)} \end{picture} \end{center} \caption{Renormalization group beta functions for the $D$-dimensional Yang-Mills gauge theory and the $d$-dimensional NLSM have the same form when $d=D-2$. (a) $D=4+\epsilon$ and $d=2+\epsilon$, (b) $D=4$ and $d=2$.} \label{fig:betafunc} \end{figure} \subsection{Spontaneous breakdown of hidden supersymmetry} \par A step of dimensional reduction is a little bit subtle. In the TQFT, the action is BRST exact by definition, so the partition function and the expectation value of the gauge invariant operator do not depend on the coupling constant. However, the NLSM$_{D-2}$ obtained after the dimensional reduction from the TQFT$_D$ is not topological and may depend on the coupling constant. This seems at first glance inconsistent. \par This problem will be resolved as follows. The dimensional reduction is a consequence of the hidden supersymmetry in TQFT obtained in the MA gauge, see \cite{KondoII}. Here the supersymmetry implies the invariance under the super rotation in the superspace $(x^\mu, \theta, \bar \theta)$, i.e., the orthsymplectic group $OSp(D|2)$. The advantage of introducing the superspace is to give a geometric meaning to the BRST transformation. In fact, the BRST symmetry becomes the translational invariance in the superspace. The BRST charges $Q_B$ and $\bar Q_B$ are generators of the translations in the direction of the Grassmann variables $\theta$ and $\bar \theta$, i.e., they are identified as ${\partial \over \partial \theta}$ and ${\partial \over \partial \bar \theta}$, respectively. In the process of the dimensional reduction we can choose arbitrary two dimensions ${\bf R}^2$ from D dimensions $(x^1, \cdots, x^D)\in {\bf R}^D$, since there is no privileged direction. However, once we have chosen specific two dimensions, the rotation symmetry is partially broken by this procedure. In this sense, the dimensional reduction causes the spontaneous breakdown of the supersymmetry hidden in TQFT. \par This becomes more clear in evaluating the expectation value of an operator based on the dimensional reduction. For this strategy to work, the support of all the operators must be contained in the (D-2) dimensional subspace to which the dimensional reduction occurs. Such an expectation value is obtained from the generating functional by restricting the external source ${\cal J}(x,\theta,\bar \theta)$ to a (D-2)-dimensional subspace ${\cal J}((z,\hat x=0),\theta=0,\bar \theta=0)$ i.e., by putting $\hat x=\theta=\bar \theta=0$. This is obtained as a consequence of restricting the orthosymplectic $OSp(D|2)$ rotation to the orthogonal $O(D-2)$ rotation, see section IV of the paper \cite{KondoII}. Of course, when a pair of quark and anti-quark exists, it is convenient to choose the (D-2)-dimensional subspace so that their trajectories are contained in the subspace when $D\ge 4$. The supersymmetry (i.e., rotational invariance $OSp(D|2)$ in the BRST superspace) is broken into the orthogonal $O(D-2)$ rotation, whereas the BRST symmetries (i.e., translational invariances in the direction of $\theta, \bar \theta$ in the superspace) is broken by putting $\hat x=\theta=\bar \theta=0$. Note that the Hilbert space of the (D-2)-dimensional bosonic theory is different from the original D-dimensional supersymmetric theory. Consequently, (D-2)-dimensional bosonic theory is no longer topological. Thus, the NLSM$_2$ can be obtained without contradiction from TQFT$_4$ by dimensional reduction. \par Finally we consider the above result from a different point of view. We consider the $4+\epsilon$ dimensional Yang-Mills theory and $2+\epsilon$ dimensional NLSM. For $\epsilon>0$, both theories have two phases, the disordered (high-temperature) phase in the strong coupling region $g>g_c$ and the ordered (low-temperature) phase in the weak coupling region $g<g_c$ where $g_c \sim O(\epsilon)$. The beta function is expected to be positive for $0<g<g_c$ and negative for $g>g_c$. See Fig.~\ref{fig:betafunc}(a). As $\epsilon$ decreases, the ordered phases shrinks and finally disappears ($g_c(\epsilon) \downarrow 0$ as $\epsilon \downarrow 0$). In this limit, the beta function becomes negative for any value of $g$, leading to the asymptotic freedom for YM$_4$ and NLSM$_2$. See Fig.~\ref{fig:betafunc}(b). In this limit, the massless Nambu-Goldstone particle associated to the spontaneous breaking of $G$ to $H$ also disappear, as examined by Bardeen, Lee and Shrock \cite{BLS76}. For $D=4$, thus, the massless Nambu-Goldstone (NG) particle associated with the spontaneous breaking of supersymmetry, if any, can not exist in two dimensions. \footnote{ The above argument is clearly insufficient in the following sense. The BRST symmetry is a global continuous symmetry. If it is spontaneously broken, the NG particle should appear, otherwise the Higgs mechanism should occur. It has been shown that the massless NG boson in question does not appear in the physical state and that the dynamical Higgs mechanism does occur. As a result, the off-diagonal gluons and the off-diagonal ghosts (anti-ghosts) become massive. See the recent paper \cite{KS00} and references cited there. } \section{Geometric meaning of gauge fixing term} \setcounter{equation}{0} In quantizing the gauge theory, the procedure of gauge fixing is indispensable to avoid infinities due to overcounting of gauge equivalent configurations. So in the quantized gauge theory we must treat the gauge fixing term seriously as well as the gauge field action. Already at the level of classical theory, it is well known that the gauge theory has a geometric meaning, i.e., gauge theory is nothing but the geometry of connection. In this section we want to emphasize that the gauge fixing term may have a geometric meaning from a viewpoint of global topology. \subsection{FP determinant} The usual procedure of gauge fixing is to insert the identity \begin{eqnarray} 1 = \Delta_{FP}[{\cal A}] \int [dU] \prod_{x} \delta (F^a[{\cal A}^U]) \end{eqnarray} into the functional integral \begin{eqnarray} Z = \int [d{\cal A}^U] \exp (- S_{YM}[{\cal A}^U]) . \end{eqnarray} Then we obtain \begin{eqnarray} Z = \int [dU] \int [d{\cal A}^U] \Delta_{FP}[{\cal A}^U] \prod_{x} \delta (F^a[{\cal A}^U]) \exp (- S_{YM}[{\cal A}^U]) , \label{Za} \end{eqnarray} since $\Delta_{FP}$ is gauge invariant, $\Delta_{FP}[{\cal A}] =\Delta_{FP}[{\cal A}^U]$. The $\Delta_{FP}$ is calculated as follows. \begin{eqnarray} \Delta_{FP}[{\cal A}]^{-1} &=& \int [d\omega] \prod_{x} \delta (F^a[{\cal A}^\omega]) \nonumber\\ &=& \int [d\omega] \sum_{k} {\delta \left( \omega - \omega_k \right) \over | \det \left( {\delta F^a[{\cal A}^\omega] \over \delta \omega} \right) |_{\omega=\omega_k} } \nonumber\\ &=& \sum_{k} {1 \over | \det \left( {\delta F^a[{\cal A}^\omega] \over \delta \omega} \right) |_{\omega=\omega_k} } . \end{eqnarray} When this result in the presence of Gribov copies is substituted into (\ref{Za}), the BRST formulation does not work. Even when there is no Gribov copies, we have the absolute value of the determinant, \begin{eqnarray} \Delta_{FP}[{\cal A}] = \Big| \det \left( {\delta F^a[{\cal A}^\omega] \over \delta \omega} \right) \Big|_{\omega=\omega_k} . \end{eqnarray} This expression is difficult to be used. Therefore, we do not adopt this approach. Rather we start from the expression, \begin{eqnarray} Z = \int [dU] \int [d{\cal A}^U] \prod_{x} \delta (F^a[{\cal A}^U]) \det \left( {\delta F^a[{\cal A}^\omega] \over \delta \omega} \right) \exp (- S_{YM}[{\cal A}^U]) . \end{eqnarray} Such a formulation was proposed by Fujikawa \cite{Fujikawa79}. We will show that such a proposal is very natural from the viewpoint of global topology. \subsection{Gauge fixing and global topology} \begin{figure} \begin{center} \unitlength=1cm \begin{picture}(13,6) \thicklines \put(3,6.4){\circle*{0.2}} \put(3,1.6){\circle*{0.2}} \bezier{300}(1,5)(3,7.8)(5,5) \bezier{300}(1,5)(0.5,4)(1,3) \bezier{300}(5,5)(5.5,4)(5,3) \bezier{300}(1,3)(3,0.2)(5,3) \thinlines \bezier{300}(0.8,4)(3,5)(5.2,4) \thicklines \bezier{300}(0.8,4)(3,3)(5.2,4) \put(3,-0.5){\bf (a)} \put(6.5,4){\Large \bf+} \put(10,6.4){\circle*{0.2}} \put(10,5.0){\circle*{0.2}} \put(10,3.0){\circle*{0.2}} \put(10,1.6){\circle*{0.2}} \bezier{300}(8,5)(10,7.8)(12,5) \bezier{300}(8,5)(7.5,4)(8,3) \bezier{300}(12,5)(12.5,4)(12,3) \bezier{300}(8,3)(10,0.2)(12,3) \put(10,4){\oval(1,2)} \thinlines \bezier{100}(7.8,4)(8.6,4.5)(9.45,4) \bezier{300}(10.5,4)(11.3,4.5)(12.2,4) \thicklines \bezier{300}(7.8,4)(8.6,3.5)(9.45,4) \bezier{300}(10.5,4)(11.3,3.5)(12.2,4) \put(10,-0.5){\bf (b)} \put(13,4){\Large \bf+ $\cdots$} \end{picture} \end{center} \caption{Various two-dimensional surfaces with different topology. The Euler number $\chi$ of the two-dimensional surface is determined by the topology of the surface, i.e., $\chi=2-2g$ for the surface with the genus $g$. The Morse index $\nu_p=(-1)^{\lambda_p}$ is a local quantity which is determined at the critical points (black dots in the figures) of the Morse function. The Euler number is equal to the sum of the Morse indices over all the critical points, i.e., $\chi=n_0-n_1+n_2$. For example, it is easy to see that (a) $\chi=2$ for the sphere $S^2$, (b) $\chi=0$ for the torus $T^2$. } \label{fig:2Dsurface} \end{figure} \par In the following, we use the notation $\phi^a = \{ {\cal A}_\mu^A(x) \}$ where $a$ denotes collectively $x,\mu,A$. If we take the gauge fixing condition \begin{eqnarray} F^a[\phi] := {\delta f(\phi) \over \delta \phi^a} = 0, \end{eqnarray} the partition function in the gauge theory is given by \begin{eqnarray} Z &:=& \int [d\phi] \prod_{a} \delta (F^a[\phi]) \det \left( {\delta F^a[\phi] \over \delta \phi^b} \right) e^{-S[\phi]} \nonumber\\ &=& \int [d\phi] \prod_{a} \delta \left( {\delta f(\phi) \over \delta \phi^a} \right) \det \left( {\delta^2 f(\phi) \over \delta \phi^a \delta \phi^b} \right) e^{-S[\phi]} . \end{eqnarray} Let $M$ be the manifold $M$ of field configurations $\{ \phi^a \}$ and $f$ a continuous function from $M$ to ${\bf R}$, $f:M \rightarrow {\bf R}$. In order to see the geometric meaning of the gauge fixing term, we consider the limit $S[\phi] \rightarrow 0$, i.e., only the gauge fixing and the corresponding FP ghost term, \begin{eqnarray} \chi_G := \int [d\phi] \prod_{a} \delta \left( {\delta f(\phi) \over \delta \phi^a} \right) \det \left( {\delta^2 f(\phi) \over \delta \phi^a \delta \phi^b} \right) . \end{eqnarray} Hence, from the property of the Dirac delta function, we have \begin{eqnarray} \chi_G &=& \int [d\phi] \sum_{k} {\delta \left( \phi - \phi_k \right) \over | \det \left( {\delta^2 f(\phi) \over \delta \phi^a \delta \phi^b} \right) |_{\phi=\phi_k} } \det \left( {\delta^2 f(\phi) \over \delta \phi^a \delta \phi^b} \right) \\ &=& \sum_{k: \nabla f(\phi_k)=0} {\rm sign} \left[ \det \left( {\delta^2 f(\phi) \over \delta \phi^a \delta \phi^b} \right)_{\phi=\phi_k} \right] , \end{eqnarray} where $\phi_k$ is a solution of $\nabla f(\phi)=0$ or $F^a[\phi]=0$. Thus we obtain \begin{eqnarray} \chi_G = \sum_{p: \nabla f(\phi_p)=0} \nu_p , \quad \nu_p := {\rm sign} (\det H_f) = (- 1)^{\lambda_p} , \label{Local} \end{eqnarray} where $H_f$ is the Hessian defined by \begin{eqnarray} H_f := {\delta^2 f(\phi) \over \delta \phi^a \delta \phi^b} . \end{eqnarray} For a smooth function $f$, the Hessian is a symmetric matrix and hence its eigenvalues are all real. The index $\lambda_P$ is equal to the number of negative eigenvalues of the Hessian. Note that $\chi$ is an integer. In order to obtain this simple expression, the existence of the FP determinant is indispensable. The function $f$ is called the Morse function \cite{Milnor63}, if all the critical points $P$ (i.e., $\nabla f(P)=0$) of $f$ are non-degenerate, i.e., $\det(H_f)\not=0$. For a finite dimensional case, it is known that all non-degenerate critical points of $f$ are isolated critical points, i.e., there is no other critical point in the neighborhood of a critical point. Whether the critical point is degenerate or non-degenerate does not depend on how to choose the coordinate systems in the manifold $M$ of field configurations $\{ \phi^a \}$. A convenient way to see the above situation is to use the standard form. The quantity $\chi$ is obtained as the sum of the index $\nu_p$ over all the critical points, once the function $f$ is given. See Fig.~\ref{fig:2Dsurface} for a simple case of two-dimensional surface $M_2$. In that case, we have \begin{eqnarray} \chi(M_2) = n_0 - n_1 + n_2 , \label{Euler1} \end{eqnarray} where $n_0, n_1, n_2$ are the total numbers of the minimal, saddle and maximum points respectively. \par For a given field configuration $\phi^a$, we can consider the global topology. The measure $[d\phi(x)]$ includes various field configurations with various global topology, each of which is characterized by an appropriate topological invariant. By repeating the similar calculation, we obtain for the partition function, \begin{eqnarray} Z = \sum_{p: \nabla f(\phi_p)=0} \nu_p e^{-S[\phi_p]} = \sum_{p: \nabla f(\phi_p)=0} (- 1)^{\lambda_p} e^{-S[\phi_p]} . \label{Z} \end{eqnarray} \par The two-dimensional case is rather simple, see Fig.~\ref{fig:2Dsurface}. The Poincare'-Hopf theorem for the two-dimensional surface states that $\chi$ defined in (\ref{Local}) is equal to the Euler number $\chi$ of the two-dimensional surface $M_2$, \begin{eqnarray} \chi(M_2) = 2 - 2g , \label{Euler2} \end{eqnarray} where $g$ is the genus, i.e., number of handles. The Morse index $\nu_p$ or $\lambda_p$ is a local quantity, but, the sum $\chi$ given by (\ref{Local}) is determined only from the global topology of the surface (\ref{Euler2}) without any local information. The Poincare-Hopf theorem gives a bridge between the local geometry and global topology: \begin{eqnarray} {\rm Local} \rightarrow \sum_{p} ({\rm local~index})_p = {\rm topological~invariant} \leftarrow {\rm Global} . \end{eqnarray} This is a very important result for our purpose. Because, by deforming the surface in a continuous way, the location of the critical point change and the Morse index at the new critical point may also change, but the total sum of the Morse index is a topological invariant which is determined only by the global topology of the surface irrespective of the way of continuous deformation. \par It is well known that the two-dimensional manifold is completely classified by the genus or the Euler number. This is not the case in higher dimensions. In fact, we must treat the infinite dimensional case. Even in the infinite dimensional case, for a specific field configuration $M_\infty^\alpha$, a topological invariant $Q(M_\infty^\alpha)$ will be determined. Then it is expected that the $\chi_G$ is expressed as a sum of topological invariants, \begin{eqnarray} \chi_G = \sum_{Q} w_Q Q(M_\infty^\alpha) . \end{eqnarray} In view of this, the functional we have chosen to derive the MA gauge, \begin{eqnarray} {\cal R}[\Omega] := \int d^Dx \ {\rm tr}_{G/H} \left( {1 \over 2} \Omega_\mu(x) \Omega_\mu(x) \right) \end{eqnarray} is considered to be a Morse function. Indeed, the MA gauge condition is obtained as the gradient of the Morse function ${\cal R}$ with respect to the large gauge transformation, \begin{eqnarray} {\delta {\cal R}[\Omega^\omega] \over \delta \omega} = F[\Omega] . \end{eqnarray} The topology change in the field configurations may be caused by the large (or finite) gauge transformation allowed in the measure $[dU]$, since the measure is invariant under the global gauge rotation $U \rightarrow e^{i\omega}U$. Thus, the gauge fixing and the associated FP term when integrated out by the functional measure $[d\phi]$ can have a geometric meaning which is related to the global topology of the field configurations. The Morse function is a tool of probing the global topology allowed in the functional space of the field configurations by gathering the local information at all the critical points. The partition function of Yang-Mills theory is given by \begin{eqnarray} Z_{YM}[0] = \sum_{p: F(\Omega_p)=0} \nu_p e^{-S_{eff}[\Omega_p]} = \sum_{p: F(\Omega_p)=0} (- 1)^{\lambda_p} e^{-S_{eff}[\Omega_p]} . \label{Z0} \end{eqnarray} \par Finally, we consider how $\chi$ changes when we change the Morse function $f$. In two-dimensional case, $\chi(M_2)$ is determined by the topology of manifold $M_2$ irrespective of the choice of Morse function $f$. If this feature survives in the infinite dimensional case, we can conclude that the global topology of the Yang-Mills theory does not depend on the way of gauge fixing. \subsection{Morse function and BRST transformation} By introducing the auxiliary field $B$ and the FP ghost and anti-ghost fields, $\psi, \bar \rho$, we can write \begin{eqnarray} \chi := \int [d\phi] \prod_{a} \delta (F^a[\phi]) \det \left( {\delta F^a[\phi] \over \delta \phi^b} \right) = \int [d\phi] [dB] [d\psi] [d\bar \rho] e^{-S_{GF}[\phi]} , \end{eqnarray} where the gauge-fixing action reads \begin{eqnarray} S_{GF} = \int d^Dx \left[ {\alpha \over 2} B^2 + BF - \bar \rho F'[\phi] \psi \right] = \int d^Dx \ \delta_B \left[ \bar \rho \left( F[\phi]+{\alpha \over 2} B \right) \right] \end{eqnarray} with the nilpotent BRST transformation, \begin{eqnarray} \delta_B \phi &=& \psi, \nonumber\\ \delta_B \psi &=& 0, \\ \delta_B \bar \rho &=& B, \\ \delta_B B &=& 0 . \end{eqnarray} By eliminating the auxiliary field $B$, we have \begin{eqnarray} \chi &=& \int [d\phi] [d\psi] [d\bar \rho] e^{-S_{GF}'}, \\ S_{GF}' &=& \int d^Dx \left[ -{1 \over 2\alpha} (F[\phi])^2 - \bar \rho F'[\phi] \psi \right] , \end{eqnarray} where \begin{eqnarray} \delta_B \phi &=& \psi, \nonumber\\ \delta_B \psi &=& 0, \\ \delta_B \bar \rho &=& {F[\phi] \over \alpha} . \end{eqnarray} The critical point $F[\phi] := {\partial f(\phi) \over \partial \phi}= 0$ corresponds to the fixed point of BRST transformation. Therefore, the integration $\int [d\phi] \delta(F[\phi]) \cdots$ is localized on the fixed point of BRST transformation. This is a characteristic feature of topological quantum field theory. The condition $F^a[\phi]=0$ is regarded as a non-linear partial differential equation. The space of parameters characterizing the solution of this equation is called the moduli space. In the TQFT, the above argument shows that the infinite dimensional functional integral reduces to finite dimensional integral on moduli space. For example, for the Yang-Mills instanton with $Q=k$, ${\rm dim}M=8k < \infty$. \section{Conclusion and discussion} \setcounter{equation}{0} In this paper we have derived a reformulation of the Yang-Mills theory based on the background field method. The reformulation identifies the Yang-Mills theory as a deformation of a topological quantum field theory as proposed in \cite{KondoII}. The background field is given by a topological soliton. \par In order to show quark confinement, the condensation of a topological soliton is necessary to occur. This has been actually derived by summing up the topological soliton contributions, provided that the topological soliton is described by the topological field theory. The topological field theory has been derived from the gauge fixing term corresponding to the {\it nonlinear} gauge fixing condition, the maximal Abelian gauge. The maximal Abelian gauge implies that the topological soliton in question is nothing but the magnetic monopole current, the four-dimensional version of the magnetic monopole. The result ensures that the quark confinement is realized in the QCD vacuum as a dual superconductor. Furthermore, we have proposed a numerical simulation which is able to confirm the validity of the above reformulation. \par We have discussed a novel mechanism for the mass generation for the gauge field, i.e., dynamical mass generation as the dimensional reduction which causes the spontaneous breakdown of the hidden supersymmetry in the topological field theory. Moreover, we have suggested that the gauge fixing action may have the geometric meaning from the view point of global topology by making use of the Morse function. \par In this paper we have restricted our consideration to the maximal Abelian gauge where the residual gauge group $H$ is the maximal torus group of the non-Abelian gauge group $G$ ($H=U(1)^{N-1}$ for $G=SU(N)$), although our formulation can be applied to any choice of $H$. Therefore the topological soliton is given by the Abelian magnetic monopole. However, it is possible to consider other choices for the residual gauge group $H$ (especially for $G=SU(N) (N \ge 3)$) which leads to the topological soliton other than the Abelian magnetic monopole, e.g., non-Abelian magnetic monopole, center vortex. Either choice will lead to the quark confinement. From the viewpoint of {\it color} confinement, however, the maximal torus group for $H$ is not necessarily the best choice. The details will be given in the subsequent paper \cite{Kondo99}.
\section{Introduction} Two observations of \object{M\,67} with the ROSAT PSPC resulted in the detection of X-ray emission from 25 members of this old open cluster (Belloni et al.\ 1993, 1998). \nocite{bellea} \nocite{bellea93} The X-ray emission of many of these sources is readily understood. For example, the X-ray emission originates in deep, hot atmospheric layers in a hot white dwarf; is due to mass transfer in a cataclysmic variable; and is caused by magnetic activity in two contact binaries and several RS CVn-type binaries. However, Belloni et al.\ (1998) point out several X-ray sources in \object{M$\,$67} for which the X-ray emission is unexplained. All but one of these objects are located away from the isochrone formed by the main sequence and the (sub)giant branch of \object{M$\,$67} (Fig.~\ref{cmd}). In this paper we investigate the nature of the X-ray emission of these stars through low- and high-resolution optical spectra. In particular, we investigate whether the emission could be coronal as a consequence of magnetic activity, by looking for emission cores in the \ion{Ca}{ii} H\&K lines. Tidal interaction in a close binary orbit is thought to enhance magnetic activity at the stellar surface by spinning up the stars in the binary. Therefore, we also derive projected rotational velocities with the crosscorrelation method. Finally, we study the H$\alpha$ profile as a possible indicator of activity or mass transfer. The observations and the data reduction are described in Sect.~2, and the analysis of the spectra in Sect.~3. Comparison with chromospherically active binaries is made in Sect.~4. A discussion of our results is given in Sect.~5. In the remainder of the introduction we give brief sketches of the stars studied in this paper; details on many of them are given by Mathieu et al.\ (1990). The stars are indicated with their number in Sanders (1977), and are listed in Table~\ref{tab1}. \nocite{mathlathea} \nocite{san} \begin{table} \caption{Stars of \object{M$\,$67} discussed in this paper. Visual magnitude and colour (from Montgomery et al.\ 1993), spectral type (from Allen \&\ Strom 1995, and Zhilinskii \&\ Frolov 1994), orbital period and eccentricity (from Mathieu et al. 1990, Latham et al.\ 1992 and -- for S\,1113 -- Mathieu et al.\ 1999, in preparation), and X-ray countrate in PSPC channels 41--240, corresponding to 0.4--2.4\,keV (from Belloni et al.\ 1998). \label{tab1}} \begin{tabular}{l@{\hspace{0.25cm}}l@{\hspace{0.25cm}}l@{\hspace{0.25cm}}l@{\hspace{0.12cm}}l@{\hspace{0.25cm}}l@{\hspace{0.25cm}}l} ID & $V$ & $B$-$V$ & sp.type & $P_{\mathrm{b}}$ & $e$ & ctrate \\ & & & & (d) & & (s$^{-1}$) \\ \\ \hline \\ S\,1040 & 11.52 & 0.87 & G4III & 42.83 & 0.027(28) & 0.0050(6) \\ S\,1063 & 13.79 & 1.05 & G8IV & 18.39 & 0.217(14) & 0.0047(6) \\ S\,1072 & 11.32 & 0.61 & G3III--IV & 1495. & 0.32(7) & 0.0013(3) \\ S\,1082 & 11.25 & 0.42 & F5IV & & & 0.0046(6) \\ S\,1113 & 13.77 & 1.01 & & 2.823 & 0.031(14) & 0.0047(6) \\ S\,1237 & 10.78 & 0.94 & G8III & 697.8 & 0.105(15) & 0.0010(3) \\ S\,1242 & 12.72 & 0.68 & & 31.78 & 0.664(18) & 0.0007(2) \\ \end{tabular} \end{table} \nocite{montea} \nocite{lathmathea} \nocite{allestro} \nocite{zhilfrol} \nocite{mathlathea} S\,1063 and S\,1113 are two binaries located below the subgiant branch in the colour-magnitude diagram of \object{M$\,$67}. Their orbital periods, 18.4 and 2.82 days respectively, are too long for them to be contact binaries; also they are too far above the main sequence to be binaries of main-sequence stars. In principle, a (sub)giant can become underluminous when it transfers mass to its companion, as energy is taken from the stellar luminosity to restore hydrostatic equilibrium (e.g.\ Kippenhahn \&\ Weigert 1967). However, mass transfer through Roche lobe overflow very rapidly leads to circularization of the binary orbit, whereas S\,1063 has an eccentricity $e=0.217$. The orbit of S\,1113 is circular, so mass transfer could be occurring in that system. For the moment, the nature of these binaries is not understood. In both, Pasquini \&\ Belloni (1998) observed emission cores in the \ion{Ca}{ii} H\&K lines. S\,1063 is reported to be photometrically variable with $\sim$ 0.10 mag (Rajamohan et al.\ 1988; Kaluzny \& Radczynska 1991), but no period is found. For S\,1113, photometric variability with a period of 0.313 days and a total amplitude of 0.6 mag was claimed by Kurochkin (1960), but this has not been confirmed by Kaluzny \& Radczynska (1991), who find variability with only 0.05 mag. S\,1063 is the only \object{M\,67} star in our sample that shows significantly variable X-ray emission (between 0.0081 and 0.0047 cts s$^{-1}$; Belloni et al.\ 1998). \nocite{kippweig} \nocite{rajaea} \nocite{kuro} \nocite{kara} \nocite{pasqbell} S\,1072 and S\,1237 are binaries with orbital periods of 1495 and 698 days, and with eccentricities $e=0.32$ and 0.105, respectively. The colour and magnitude of S\,1072 cannot be explained with the pairing of a giant and a blue straggler, since this is not compatible with its $ubvy$ photometry (Nissen et al.\ 1987; Mathieu \&\ Latham 1986), nor with superposition of three subgiants, since this is excluded by the radial velocity correlations (Mathieu et al.\ 1990). The absence of the 6708 \AA\ lithium feature in the spectrum of S\,1072 indicates that the surface material has undergone mixing (Hobbs \&\ Mathieu 1991; Pritchett \& Glaspey 1991). S\,1237 could be a binary of a giant and a star at the top of the evolved main sequence (Janes \& Smith 1984); high-resolution spectroscopy should be able to detect the main-sequence star in that case (Mathieu et al.\ 1990). The wide orbits and significant eccentricities appear to exclude both mass transfer and tidal interaction as explanations for the X-ray emission. \nocite{nissea} \nocite{hobbmath} \nocite{pritglas} \nocite{janesmit} \nocite{mathlath} S\,1242 has the largest eccentricity of the binaries in our sample, at $e=0.66$ in an orbit of 31.8$\,$days. Its position on the subgiant branch is explained if a subgiant of 1.25\,$M_{\sun}$ has a secondary with $V>15$ (Mathieu et al.\ 1990). \ion{Ca}{ii} K line emission is reported by Pasquini \&\ Belloni (1998). Photometric variability with a period of 4.88$\,$days and amplitude of 0.0025 mag has been found by Gilliland et al.\ (1991). We note that this photometric period corresponds to corotation with the orbit at periastron, which suggests that the X-ray emission may be due to tidal interaction taking place at periastron. The binary would then be an interesting example of a system in transition from an eccentric to a circular orbit. Indeed, according to the diagnostic diagram of Verbunt \&\ Phinney (1995) a giant of 1.25\,$M_{\sun}$ with a current radius of $\simeq2.3\,R_{\sun}$ (as derived from the location of S\,1242 in the colour-magnitude diagram) cannot have circularized an orbit of 31.8\,days. \nocite{gillea} \nocite{verbphin} S\,1040 is a binary consisting of a giant and a white dwarf. The progenitor of the white dwarf circularized the orbit during a phase of mass transfer (Verbunt \&\ Phinney 1995); as a result the mass of the white dwarf is very low (Landsman et al.\ 1997). The white dwarf is probably too cool, at 16\,160$\,$K, to be the X-ray emitter. Indications for magnetic activity are \ion{Ca}{ii} H\&K (Pasquini \&\ Belloni 1998) and \ion{Mg}{ii} ($\lambda\lambda$ 2800 \AA, Landsman et al.\ 1997) emission lines. If the X-rays are due to coronal emission of the giant, this must be the consequence of the past evolution of the binary, since the giant is too small for significant tidal interaction to be taking place in the current orbit. \nocite{landea} S\,1082 is a blue straggler. Photometric variability of 0.08 mag within a few hours was observed by Simoda (1991). Goranskii et al.\ (1992) found eclipses with a total amplitude of 0.12 mag and a binary period of 1.07 days; however, the radial velocities of the star do not show this period, and vary by about 2\,km s$^{-1}$, far too little for a 1\,day eclipsing binary (Mathieu et al.\ 1986). Landsman et al.\ (1998) detect a significant excess at 1520 \AA\ with the Ultraviolet Imaging Telescope, and ascribe this to a hot, subluminous secondary. Such a secondary was suggested already by Mathys (1991) on the basis of a broad component in the \ion{Na}{i} D and \ion{O}{i} absorption lines. \nocite{goraea} \nocite{mathlath} \nocite{math} \nocite{landea} \nocite{simo} \nocite{mathea86} \begin{figure} \centerline{\psfig{figure=M67.ps,angle=-90,width=\columnwidth,clip=t} {\hfil}} \caption{Colour-magnitude diagram of \object{M$\,$67}. Colours and magnitudes are from Montgomery et al.\ (1993). Only stars with membership probability $> 0.8$ (based on their proper motion, Sanders 1977) were selected. Stars indicate the observed X-ray binaries; squares two member giants observed for comparison. } \label{cmd} \end{figure} \nocite{montea} \nocite{san} \section{Observations and data reduction} Optical spectra were obtained on February 28/29, 1996 with the 4.2m William Herschel Telescope on La Palma, under good weather conditions (seeing $<1\arcsec$ until 4$\fh$30 UT, $<2\arcsec$ thereafter). In addition to the X-ray sources in \object{M$\,$67} we observed two ordinary member giants of \object{M$\,$67}, S\,1288 and S\,1402, for comparison. Furthermore one flux standard and three velocity standards were observed. The blue high-resolution spectra of S\,1113 were obtained on April 7/8, 1998 with the same telescope through a service observation (seeing 1--2$\arcsec$). A log of the observations is given in Table~\ref{tab11}. \begin{table} \caption{Log of the observations. For each target we give the UT at start of the exposures and the exposure time for the ISIS and UES blue and red spectra. All observations were obtained on 28/29 February 1996, except the blue UES exposures of S\,1113 which were taken on 7/8 April 1998 (indicated with $^*$ in the table). \label{tab11}} \begin{tabular}{llrlrlr} ID & \multicolumn{2}{l}{ISIS} & \multicolumn{2}{l}{UES blue} & \multicolumn{2}{l}{UES red} \\ & \multicolumn{2}{l}{---------------} & \multicolumn{2}{l}{---------------} & \multicolumn{2}{l}{---------------} \\ & UT & \multicolumn{1}{l}{$t_{\mathrm{exp}}$}& UT & \multicolumn{1}{l}{$t_{\mathrm{exp}}$} & UT & \multicolumn{1}{l}{$t_{\mathrm{exp}}$} \\ & & \multicolumn{1}{l}{(s)} & & \multicolumn{1}{l}{(s)} & & \multicolumn{1}{l}{(s)} \\ \\ \hline \\ \multicolumn{5}{l}{{\sc targets}} \\ S\,1040 & 02:43&60 & 22:03&600 & 00:46&300 \\ S\,1063 & 02:30&180 & 20:51&1200 & 00:11&1200 \\ & & & 21:15&1200 & 00:33&600 \\ & & & 21:37&1200\\ S\,1072 & 02:49&60 & 22:24&360 & 01:01&240 \\ S\,1082 & 02:39&60 & 20:39&300 & 00:01&360 \\ & & & 23:24&600 & 01:41&360 \\ & & & & & 04:11&360 \\ S\,1113 & 03:01&180 & 22:37$^*$ & 900 & 03:36&1200 \\ & & & 22:58$^*$ & 900 & 03:58&600 \\ S\,1237 & 02:46&60 & 22:16&300 & 00:54&180 \\ S\,1242 & 02:55&120 & 22:33&1500 & 01:08&900 \\ \\ \multicolumn{5}{l}{{\sc comparison giants}} \\ S\,1288 & 02:58&60 & 23:12&600 & 01:33&300 \\ S\,1402 & 02:52&60 & 23:01&450 & 01:26&240 \\ \\ \multicolumn{5}{l}{{\sc flux standard}} \\ \object{HZ\,44} & 02:21&80 & 23:42&600 & 01:57&600 \\ & 03:13&80 \\ & 04:43&480 \\ \\ \multicolumn{5}{l}{{\sc velocity standards}} \\ \object{HD$\,$132737} & & & 06:04& 90 & 05:13&45 \\ \object{HD$\,$136202} & & & 05:55&100 & 05:42&50 \\ \object{HD$\,$171232} & & & & & 05:34&45 \\ \end{tabular} \end{table} All spectra have been reduced using the Image Reduction and Analysis Facility (IRAF)\footnote{IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the National Science Foundation}. \subsection{Low-resolution spectra} Low-resolution spectra were taken with the ISIS double-beam spectrograph (Carter et al.\ 1993). The blue arm of ISIS was used with the 300 lines per mm grating and TEK-CCD, resulting in a wavelength coverage of 3831 to 5404 \AA\ and a dispersion of 1.54 \AA\ per pixel at 4000 \AA. The red arm, combined with the 316 lines per mm grating and EEV-CCD, covered a wavelength region of 5619 to 7135 \AA\ with a dispersion of 1.40 \AA\ per pixel at 6500 \AA. The format of the frames is 1124 $\times$ 200 pixels which includes the under- and overscan regions. For the object exposures the slit width was set to $4\arcsec$. Flatfields were made with a Tungsten lamp while CuAr and CuNe lamp exposures were taken for the purpose of wavelength calibration. \nocite{cartea} For the ISIS-spectra, basic reduction steps have been done within the IRAF {\sc ccdred}-package. These steps include removing the bias signal making use of the under- and overscan regions and zero frames, trimming the frames to remove the under- and overscan, and flatfielding to correct for small pixel-to-pixel gain variations. The remaining reduction has been done with IRAF {\sc specred}-package tasks. With the optimal extraction algorithm (Horne 1986) the two dimensional images are reduced to one dimensional spectra. Next, the spectra are calibrated in wavelength with the arc frames. A dispersion solution is found by fitting third (blue) and fourth (red) order polynomials to the positions on the CCD of the arclamp lines. The fluxes of the spectra are calibrated with the absolute fluxes of HZ\,44, tabulated at 50 \AA\ intervals (Massey et al. 1988), and adopting the standard atmospheric extinction curve for La Palma as given by King (1985). The estimated accuracy of the flux calibration is $\sim 10$\%. \nocite{horn}\nocite{king} \nocite{massea} \subsection{High-resolution spectra} High-resolution echelle spectra were taken with~the Utrecht Echelle Spectrograph (UES, Unger et al.\ 1993). Observations were done with a $\sim 1\arcsec$ slitwidth. \nocite{ungeea} For the 1996 observations, the UES was used in combination with a 1024 $\times$ 1024 pixels TEK-CCD, and the 31.6 lines per mm grating (E31), which resulted in a broad wavelength coverage, but small separation of the echelle orders on the CCD. In this setup, the UES resolving power is 49\,000 per resolution element (two pixels), corresponding to a dispersion of 3 km s$^{-1}$ per pixel or 0.06 \AA\ per pixel at 6000 \AA. The frames were centered on $\lambda_\mathrm{cen}=4250$ \AA\ and $\lambda_\mathrm{cen}=5930$ \AA\ in order to get a blue (3820 to 4920 \AA) and red (4890 to 7940 \AA) echelle spectrum. The number of orders recorded on the CCD is 34 in the blue and 45 in the red, each covering $\sim$ 45 to 80 \AA\ increasing for longer wavelengths. Towards the red, gaps occur between the wavelength coverage of adjacent orders. Exposures of a quartz lamp were taken to make the flatfield corrections. ThAr exposures served as wavelength calibration frames. For the 1998 observations of S\,1113, a 2048 $\times$ 2048 pixels SITe-CCD was used. Two spectra were taken with the 79.0 lines per mm (E79) grating ($\lambda_\mathrm{cen}=4343$ \AA). The difference between the E79 and the E31 gratings is that E79-spectra have a larger separation of the echelle-orders on the detector, which can improve the determination of the sky-background. The spectral resolution of the gratings is the same (the central dispersion in these observations is $\sim 0.04$ \AA\ per pixel). The specified wavelength-coverage for this combination of grating and detector is 3546 to 6163 \AA\, but only the central orders were bright enough to extract spectra (3724 to 5998 \AA). Flatfield and ThAr exposures were made for calibration purposes. The reduction of the UES spectra has been performed using the routines available within the IRAF {\sc echelle}-package. First, the frames are debiased and the under- and overscan regions removed. After locating the orders on the CCD for both the quartz lamp and the object exposures, we flatfielded the frames. Spectra are extracted with optimal extraction. The small order separation makes sky subtraction difficult; however, our targets are bright, and the resulting error is negligible. In the step of wavelength calibration, the dispersion solution is derived by fitting third and fourth order polynomials leaving rms-residuals of 0.004 \AA\ (red) and 0.002 \AA\ (blue, 0.003 \AA\ for the 1998-spectra). To find absolute fluxes for the \ion{Ca}{ii} K ($\lambda \, 3933.67$ \AA) \& H ($\lambda \, 3968.47$ \AA) emission lines (Sect.~3.1), the fluxes of the relevant blue orders of an object have been calibrated with the calibrated ISIS spectrum of the same object. Continuum normalization of the orders in the red spectra, required for the rotational velocity analysis, is done by fitting third to fifth order polynomials to the wavelength-calibrated spectra. \section{Data analysis} We study two indicators of magnetic activity. The direct indicator is emission in the cores of the \ion{Ca}{ii} H\&K lines. Another indicator is the rotational speed: rapid (differential) rotation and convective motions are thought to generate magnetic fields through a dynamo. \subsection{Determination of \ion{Ca}{ii} H\&K emission fluxes} To estimate the amount of flux emitted in the \ion{Ca}{ii} H\&K line cores, $F_{\mathrm{Ca}}$, we add the fluxes above the H\&K absorption profiles as follows. An upper and a lower limit of the level of the absorption pseudo-continuum is estimated by eye and is marked by a straight line. For S\,1113 this is illustrated in Fig.~\ref{subca}. We obtain a lower and upper limit of the emitted flux by adding the fluxes in each wavelength-bin above these levels. The value given in Table~\ref{fluxes} is the average of these two results, the uncertainty is half their difference. Use of higher order fits (following Fern\'andez-Figueroa et al.\ 1994) to the absorption profile gives similar results. If an emission line is not clearly visible, we obtain an upper limit by estimating the minimal detectable emission flux at the H\&K line centers within a 1 \AA\ wide region (typical width of the emission lines). Six of our sources show \ion{Ca}{ii} H\&K line emission (Fig.~\ref{cahk}). The profiles of S\,1113 appear to be double-lined, suggesting that we see activity of both stars (Fig.~\ref{subca}). The fluxes given in Table~\ref{fluxes} are the total fluxes, i.e.\ no attempt was made to deblend the emission lines. No emission is visible in the spectrum of S\,1082. \begin{figure} \centerline{\psfig{figure=plotca.ps,width=\columnwidth,clip=t} {\hfil}} \caption{The \ion{Ca}{ii} H\&K regions in the high-resolution spectra of our targets. In the lower right corner the spectrum of the non-active comparison giant S\,1288 is shown.} \label{cahk} \end{figure} \begin{table} \caption{Fluxes of emission cores in \ion{Ca}{ii} H\&K lines and projected rotational velocities derived from our high-resolution spectra. The velocities were determined from crosscorrelation with the spectrum of \object{HD\,136202} (F8III-IV, ESA 1997) for S\,1040, S\,1072, S\,1082 and S\,1242; and with the spectrum of \object{HD\,171232} (G8III, ESA 1997) for the other stars. Both HD stars have line widths $\tau=1.8\pm$0.1\,km s$^{-1}$. For S\,1113 we list both components; the primary contributes $\sim 82$\%\ of the light and has broader lines. \label{fluxes}} \begin{tabular}{lrrl} ID & \multicolumn{1}{l}{$\log F_{\ion{Ca}{k}}$} & \multicolumn{1}{l}{$\log F_{\ion{Ca}{h}}$} & \multicolumn{1}{l}{$v\sin i$} \\ & \multicolumn{1}{l}{(${\rm erg}\,{\rm cm}^{-2}\,{\rm s}^{-1}$)} & \multicolumn{1}{l}{(${\rm erg}\,{\rm cm}^{-2}\,{\rm s}^{-1}$)} & \multicolumn{1}{l}{(km s$^{-1}$)} \\ \\ \hline \\ S\,1040 & $16(4)\times10^{-15}$ & $17(5)\times10^{-15}$ & 3.0(1.0) \\ S\,1063 & $6(1)\times10^{-15}$ & $7(1)\times10^{-15}$ & 3.9(0.8) \\ S\,1072 & $3(2)\times10^{-15}$ & $<13\times10^{-15}$ & 8.1(1.1) \\ S\,1082 & $ <11\times10^{-15}$ & $<24\times10^{-15}$ & 5.1(0.7) \\ S\,1113 & $9(2)\times10^{-15}$ & $10(3)\times10^{-15}$ & 45(6) \\ & & & 12(1) \\ S\,1237 & $4(2)\times10^{-15}$ & $8(5)\times10^{-15}$ & $<1.8$ \\ S\,1242 & $2.2(0.5)\times10^{-15}$ & $2(1)\times10^{-15}$ & $<2.6(1.0)$ \\ \end{tabular} \end{table} \nocite{hipp} \subsection{Determination of projected rotational velocities} \subsubsection{Crosscorrelation} In order to derive the projected rotational velocity $v\sin i$ of our targets, we apply the crosscorrelation technique (e.g. Tonry \& Davis 1979). This method computes the correlation between the object spectrum and an appropriately chosen template spectrum as function of relative shift. The position of the maximum of the crosscorrelation function ({\sc ccf}) provides the value of the radial-velocity difference between object and template. The width of the peak is indicative for the width of the spectral lines and can therefore be used as a measure of the rotational velocity of the stars. \nocite{tonrdavi} For rotational velocities not too large the line profiles may be approximated with Gaussians, allowing analytical treatment of the crosscorrelation method. Assuming that the binary spectrum is a shifted, scaled and broadened version of the template spectrum, the broadening can be related to the width of the {\sc ccf} peak as follows (Gunn et al.\ 1996). With $\tau$ the dispersion in the template's and $\beta$ the dispersion in the target's spectral lines, $\mu$ the dispersion of the {\sc ccf} peak and $\sigma$ the dispersion of the gaussian that describes the broadening of the object's spectrum with respect to the template, one can write: \begin{equation} \mu^2=\tau^2 + \beta^2=2\tau^2 + \sigma^2 \label{sigma}\end{equation} Eq.~\ref{sigma} applies to both components in the binary spectrum and their corresponding {\sc ccf} peaks. \nocite{gunnea} The crosscorrelations are performed with the IRAF task {\sc fxcor} that uses Fourier transforms of the spectra to compute the {\sc ccf}. Before performing the crosscorrelation, the continuum is subtracted from the normalized spectra. Filtering in the Fourier domain is applied to avoid undesirable contributions originating from noise or intrinsically broad lines (see Wyatt 1985). \nocite{wyat} The templates are chosen from the radial-velocity standards such that their spectral types resemble those of the targets. The value of $\tau$ for these stars is determined for each order separately by autocorrelation of the template spectrum adopting the same filter used for the crosscorrelations. In this case $\sigma$ is zero and therefore $\tau$ is found directly from the width of the {\sc ccf} peak: ${\tau}^2={\mu}^2/2$. The template spectra are correlated with our target stars order by order, where we limit ourselves to orders in the red spectra that do not suffer from strong telluric lines. Most {\sc ccf} peaks can be fitted well with a gaussian. As the final value for $v\sin i$ we give the broadening $\sigma$ averaged over the different orders, and for the uncertainty we take the rms of the spread around the average of $\sigma$ (Table~\ref{fluxes}). Equating $v\sin i$ to $\sigma$ implicitly assumes that $\tau$ is the width of the lines not related to rotation. An upper limit to $v\sin i$ is found from the other extreme in which we assume that the total width of the spectral line $\beta$ follows from rotation. This creates uncertainties of the order of $\sigma$ for S\,1242. In the case of S\,1237 we find that $\beta < \tau$ ($\sigma < 0$), i.e. the lines in S\,1237 are narrower than in the template. For these two stars, we give an upper limit of $v\sin i < \beta$ in Table~\ref{fluxes}. S1113 is the only binary observed whose {\sc ccf} shows two peaks. The {\sc ccf} peaks of both stars overlap in the 1998 observation. Therefore we do not use these spectra in the following analysis. For the 1996 spectra, the {\sc ccf} shows two peaks, one of which is broad indicating the presence of a fast rotating star. Both peaks are clearly separated with a center-to-center velocity separation of $\sim$ 110 km s$^{-1}$. The lines in the spectrum are smeared out and less pronounced, resulting in a noisy {\sc ccf}. To improve this, we combine four sequential orders before crosscorrelating (being constrained by the maximum number of points {\sc fxcor} can handle). From eq.~17 in Gunn et al. we derive the relative light contribution of both components to the spectrum from the height and dispersion of the crosscorrelation peaks, assuming that the binary stars have the same spectrum as the template ($\alpha=1$ in eq.~17). According to this the rapidly rotating star contributes $\sim 82$\%\ of the light. Note that luminosity ratios derived from cross-correlations are uncertain and should be confirmed photometrically. \subsubsection{Fourier-Bessel transformation} The line profile of the fast rotating star in S\,1113 is not compatible anymore with a Gaussian, therefore we adopt another method to determine its $v\sin i$, described in Piters et al. (1996). This method uses the property that the Fourier-Bessel transform of a spectral line that is purely rotationally broadened has a maximum at the position of the projected rotational velocity. In practice, this position is a function of the limits over which the Fourier transform is performed. The local maxima of this velocity-versus-cutoff-frequency (vcf) function approach $v\sin i$ within half a percent, the result growing more accurate for maxima at higher frequencies. This error is negligible when compared to errors arising from noise, other line broadening mechanisms, etc. (see Piters et al. 1996). In our determination of $v\sin i$ of the primary of S\,1113, we have used the first local maximum in the vcf-plot of the transformation of four isolated \ion{Fe}{i} lines at $\lambda\lambda$ 6265.14, 6400.15, 6408.03 and 6411.54 \AA. The $v\sin i$ in Table~\ref{fluxes} is an average of the resulting values; the $v\sin i$ of the secondary is found from crosscorrelation. The application of the Fourier-Bessel transformation method is limited on the low velocity side by the spectral resolution: the Fourier transform cannot be performed beyond the Nyquist frequency which for slow rotators lies at a frequency that is lower than the cutoff-frequency at which the first maximum occurs in the vcf-plot. For our spectra this means that the method cannot be used for $v\sin i <$ 5.2 km s$^{-1}$. Indeed, for every star for which the crosscorrelation method gives a $v\sin i$ smaller than this value, the vcf-plot does not reach the first local maximum, except for S\,1082 for which we find $v\sin i$ = 9.5(1.6). For S\,1072, the Fourier-Bessel transform gives a $v\sin i$ of 12.7(1.0) km s$^{-1}$. Spectral lines were selected from those used in Groot et al. (1996) and from the additional lines used for S\,1113. \nocite{pitegrooea} \nocite{grootea} \section{Results} The results of our search for emission cores in the \ion{Ca}{ii} H\&K lines are displayed in Fig.~\ref{cahk} and in Table~\ref{fluxes}. The emission lines are strong in S\,1063, S\,1113 and S\,1040; and still detectable in S\,1242 which indicates chromospheric activity in these stars. In S\,1072 and S\,1237 the emission cores are marginal, and on S\,1082 we can only determine an upper limit. The (projected) rotational velocities of all of our stars are relatively small $v\sin i<10\,$km s$^{-1}$, with the exception of S\,1113. In Sect.~4.1 we investigate whether the relations between X-ray emission, strength of the emission cores in the \ion{Ca}{ii} H\&K lines, and the rotational velocities of the unusual X-ray emitters in \object{M$\,$67} are similar to the relations found for well-known magnetically active stars, the RS CVn binaries. In Sect.~4.2 we briefly discuss the behaviour of the H$\alpha$ line and spectral lines other than \ion{Ca}{ii} H\&K that are indicators of chromospheric activity. Individual systems are discussed in Sect.~4.3. \subsection{Comparison with RS CVn binaries} To investigate whether the X-rays of the \object{M$\,$67} stars studied in this paper are related to magnetic activity, we compare their optical activity indicators and X-ray fluxes with those of a sample of RS CVn binaries. In particular, we select RS CVn binaries for which fluxes of the emission cores in the \ion{Ca}{ii} H\&K lines have been determined from high-resolution spectra by Fern\'andez-Figueroa et al.\ (1994). To obtain X-ray countrates for these binaries, we searched the ROSAT data archive for PSPC observations of them. We then analyzed all these observations, and determined the countrates, in the same bandpass as used in the analysis of \object{M\,67}, using the standard procedure described in Zimmermann et al.\ (1994). All pointings that we have analyzed actually led to a positive detection of the RS CVn system: even when not the target of the observation, the RS CVn system is usually the brightest object in the field of view. The results of our analysis are listed in Table~\ref{tabros}. \nocite{fernea} \nocite{zimmea94} \begin{table} \caption{ROSAT PSPC countrates for RS~CVn binaries, from our analysis of archival data. For each source we list the distance, adopted from the Hipparcos catalogue (ESA 1997), the Julian date ($-$2440000) of the beginning of the ROSAT exposure, the effective exposure time, the countrate in channels 41--240 (i.e.\ roughly in the 0.4--2.4\,keV band), and the offset of the star to the ROSAT pointing direction. Where applicable we also refer to earlier publication of the ROSAT observation: $^a$Singh et al. (1996a), $^b$Welty \& Ramsey (1995), $^c$White et al. (1994), $^d$Yi et al. (1997), $^e$Singh et al. (1996b), $^f$Bauer \&\ Bregman (1996), $^g$Ortolani et al. (1997). \label{tabros}} \begin{tabular}{lllrl@{\hspace{0.05cm}}r} name & d & JD & \multicolumn{1}{l}{$t_{\mathrm{exp}}$} & ctrate & \multicolumn{1}{l}{$\Delta$} \\ & (pc) & & \multicolumn{1}{l}{(s)} & (s$^{-1}$) & \multicolumn{1}{l}{($\arcmin$)} \\ \\ \hline \\ \multicolumn{6}{l}{{\sc Luminosity class V (Group 1)}} \\ \object{IL Com} & 107$\pm$12 & 8420.575 & 16156 & 0.322(5) & 36 \\ & & 8775.123 & 8345 & 0.126(5) & 36 \\ \object{TZ CrB} & 21.7$\pm$0.5 & 8864.531 & 4267 & 6.072(15) & 0 \\ & & 9003.872 & 5776 & 6.228(11) & 0 \\ & & 8864.595 & 3651 & 6.84(2) & 24 \\ & & 9004.871 & 3151 & 5.41(2) & 24 \\ \object{V772 Her} & 37.7$\pm$1.9 & 9049.117 & 14531 & 1.206(4) & 0 \\ \object{BY Dra} & 16.4$\pm$0.2 & 9247.809 & 9895 & 1.009(7) & 0 \\ \object{V775 Her}$^a$ & 21.4$\pm$0.5 & 9085.714 & 2140 & 1.10(2) & 0 \\ \object{ER Vul} & 49.9$\pm$2.1 & 9148.099 & 1209 & 1.23(3) & 0 \\ \object{KZ And} & 25.3$\pm$4.9 & 8604.712 & 5249 & 0.888(16) & 36 \\ \\ \multicolumn{6}{l}{{\sc Luminosity class IV (Group 2)}} \\ \object{V711 Tau} & 29.0$\pm$7 & 8648.446 & 3098 & 6.29(2) & 0 \\ \object{UX Com} & 168$\pm$51 & 8426.486 & 21428 & 0.146(3) & 33 \\ \object{RS CVn} & 108$\pm$12 & 8810.279 & 2526 & 0.336(17) & 44 \\ & & 8991.040 & 6214 & 0.372(11) & 44 \\ & & 8796.874 & 5076 & 0.305(12) & 51 \\ & & 8966.506 & 2904 & 0.366(18) & 51 \\ \object{HR 5110} & 44.5$\pm$1.2 & 8431.422 & 71803 & 2.3152(14) & 44 \\ & & 9158.435 & 37658 & 2.403(3) & 44 \\ \object{SS Boo}$^b$ & 202$\pm$57 & 9030.359 & 10327 & 0.059(2) & 0 \\ \object{RT Lac}$^b$ & 193$\pm$39 & 8789.392 & 8668 & 0.198(5) & 0 \\ \object{AR Lac}$^c$ & 42.0$\pm$1.0 & 8620.767 & 13460 & 1.717(4) & 0 \\ & & 9136.873 & 4892 & 2.902(13) & 0 \\ \\ \multicolumn{6}{l}{{\sc Luminosity class III/II (Group 3)}} \\ \object{12 Cam} & 192$\pm$34 & 8322.477 & 3516 & 0.574(13) & 3 \\ \object{$\sigma$ Gem}$^d$ & 37.5$\pm$1.1 & 8346.927 & 4745 & 5.269(14) & 0 \\ & & 8904.569 & 1438 & 3.43(4) & 0 \\ & & 8735.159 & 7940 & 4.753(8) & 0 \\ \object{DK Dra} & 138$\pm$10 & 9272.484 & 4156 & 1.32(2) & 41 \\ \object{$\epsilon$ UMi} & 106$\pm$7.6 & 8328.919 & 14690 & 0.624(6) & 40 \\ \object{DR Dra}$^e$ & 103.$\pm$8.5& 8863.981 & 10576 & 1.554(8) & 31 \\ \object{HR 7428}$^e$ & 323$\pm$53 & 9088.631 & 24889 & 0.112(2) & 0 \\ \object{IM Peg} & 96.8$\pm$7.1 & 8589.200 & 6335 & 1.692(15) & 44 \\ & & 8769.630 & 8150 & 1.422(12) & 44 \\ & & 8973.263 & 22143 & 1.909(4) & 44 \\ \object{$\lambda$ And}$^{f, g}$ & 25.8$\pm$0.5 & 8448.155 & 31165 & 4.077(2) & 0 \\ \end{tabular} \end{table} \nocite{bauebreg} \nocite{singeaa} \nocite{ortoea} \nocite{whitea} \nocite{weltrams} \nocite{yielea} \nocite{singeab} To compare systems at different distances, we multiply the ROSAT countrate and the flux of the emission cores for each system with the square of the distance listed in Table~\ref{tabros}; for \object{M$\,$67} we adopt a distance of 850\,pc (Twarog \& Anthony-Twarog 1989). No corrections are made for interstellar absorption. The choice of the 0.4--2.4\,keV bandpass minimizes the effects of interstellar absorption, which are severe at energies $<0.4$\,keV. As it is unknown which component of the binary emits the X-rays, we plot the total X-ray and \ion{Ca}{ii} fluxes, adding the contributions of both components where these are given separately by Fern\'andez-Figueroa et al.\ (1994). \nocite{twan} The resulting 'absolute' countrates and fluxes are shown in Fig.~\ref{xca}. The \object{M$\,$67} systems with \ion{Ca}{ii} H\&K emission clearly visible in Fig.~\ref{cahk}, viz.\ S\,1063, S\,1113, S\,1040 and S\,1242 lie on the relation between X-ray and \ion{Ca}{ii} H\&K emission defined by the RS CVn systems, in agreement with the hypothesis that the X-ray flux of these objects is related to the magnetic activity. It is also seen that the upper limits or marginally detected emission cores in S\,1082, S\,1072 and S\,1237 are high enough that we cannot exclude the hypothesis that the X-ray emission in these systems is related to magnetic activity. The rotational velocity is another indicator of magnetic activity. We investigate the relation between rotational velocity and Ca emission by selecting those stars from the sample of Fern\'andez-Figueroa for which a value of $v\sin i$ is given in the Catalogue of Chromospherically Active Binary Stars (Strassmeier et al. 1993). In Fig.~\ref{vsinica} the \ion{Ca}{ii} H\&K emission of these stars is compared with their $v\sin i$. In this figure we do discriminate between the separate contributions of both stars to $F_{\mathrm{Ca}}$, with the exception of S\,1113 for which we combine the total flux $F_{\mathrm{Ca}}$ with the $v\sin i$ of the primary. The \object{M$\,$67} stars are found within the range occupied by chromospherically active stars. We note that the correlation between the observed H\&K flux and $v\sin i$ is not tight. In particular, high and low \ion{Ca}{ii} H\&K emission flux is found at low values of $v\sin i$. Some of the scatter may be due to the use of $v\sin i$ instead of the stellar rotation period. Parameters depending on the spectral type (e.g. properties of the convective region) have been used to reduce the scatter in the activity-rotation relation; whereas this is successful for main-sequence stars with $0.5\mathrel{\hbox{\rlap{\lower.55ex \hbox {$\sim$} B-V\mathrel{\hbox{\rlap{\lower.55ex \hbox {$\sim$} 0.8$, it fails for other main-sequence stars and for giants (see discussion in St\c{e}pie\'n 1994). For example, the three giants \object{33 Psc} (K0 III), \object{12 Cam} (K0 III) and \object{DR Dra} (K0-2 III) have $v\sin i$ values of 10, 10 and 8 km s$^{-1}$, respectively, but differ in $\log d^2 F_{\mathrm{Ca}}$ by three orders of magnitude (see Fig. \ref{vsinica}). \nocite{straea} \nocite{step} \begin{figure} \centerline{\psfig{figure=xca_41.ps,width=\columnwidth,angle=-90,clip=t} {\hfil}} \caption{PSPC countrates in channels 41--240 versus observed \ion{Ca}{ii} H\&K emission flux $F_{\mathrm{Ca}}$ (in erg s$^{-1}$cm$^{-2}$). Both are multiplied with the square of the distance (in pc). Open circles are chromospherically active binaries from the sample of Fern\'andez-Figueroa (1994). Their size indicates the luminosity class of the active component. When more than one PSPC observation is available, the countrate of the longest exposure is plotted. Filled symbols are the \object{M\,67}-sources. Triangles indicate upper limits.} \label{xca} \end{figure} \nocite{fernea} \begin{figure} \centerline{\psfig{figure=cafluxvsini.ps,angle=-90,width=\columnwidth,clip=t} {\hfil}} \caption{Observed \ion{Ca}{ii} H\&K emission flux $F_{\mathrm{Ca}}$ versus $v\sin i$. Open circles show the comparison sample of chromospherically active binaries up to 60 km s$^{-1}$ (Strassmeier et al. 1993 and Fern\'andez-Figueroa 1994). Their size indicates the luminosity class of the active star. Vertical bars indicate 1-$\sigma$ errors due to uncertainty in the distance, for systems with a Hipparcos parallax. \object{M\,67} sources are plotted as filled symbols. \label{vsinica} } \end{figure} \nocite{straea} \subsection{Activity indicators} The H$\alpha$ lines ($\lambda$ 6562.76 \AA) of only two stars, S\,1063 and S\,1113, show clear evidence of emission, as shown in Fig.~\ref{subha}. This is described in more detail in their individual subsection in Sect.~4.3. For the other \object{M$\,$67} stars, we have used the few (sub)giants in the library of UES-spectra of Montes \&\ Mart\'{\i}n (1998) to investigate the behaviour of H$\alpha$. We have chosen library spectra of stars that match the spectra of the \object{M$\,$67} stars as closely as possible (see Fig.~\ref{halpha}). S\,1072 and S\,1237 show no evidence for filling in of the H$\alpha$ profile compared to a G0IV--V and a K0III star, respectively. In S\,1040 H$\alpha$ seems slightly filled in compared to a G8IV and a K0III star. This is also the case for S\,1242 compared to a G0IV-V star, but we note that no classification for this star is found in literature. For S\,1082, no matching spectrum is available in this wavelength region. \nocite{montmart} Filling in of the lines in the \ion{Mg}{i} b triplet ($\lambda\lambda$ 5167.33, 5172.70 and 5183.62 \AA) and in the \ion{Na}{i} D doublet ($\lambda\lambda$ 5889.95 and 5895.92 \AA) is visible in some active stars. The presence of a \ion{He}{i} D$_3$ ($\lambda$ 5876.56) absorption or emission feature can also indicate activity (see discussion in Montes \&\ Mart\'{\i}n (1998) and references therein). However, in none of the \object{M$\,$67} stars we see filled in \ion{Mg}{i} b and \ion{Na}{i} D lines. Neither do we see a clear \ion{He}{i} D$_3$ feature. For S\,1082 (\ion{Mg}{i} b and \ion{Na}{i} D) and S\,1113 (\ion{Mg}{i} b, \ion{Na}{i} D and \ion{He}{i} D$_3$) we find no suitable library stars for these features. \begin{figure} \centerline{\psfig{figure=halpha.ps,width=\columnwidth,angle=-90,clip=t} {\hfil}} \caption{H$\alpha$ profiles of S\,1040, S\,1072, S\,1237 and S\,1242. Library UES-spectra (Montes \&\ Mart\'{\i}n 1998) are plotted with a thinner line. The comparison star is a K0III giant (\object{HD\,48432}) for S\,1040 and S\,1237, and a G0IV--V subgiant (\object{HD\,160269}) for the other stars. Position of H$\alpha$ and other lines are indicated with vertical lines.} \label{halpha} \end{figure} \subsection{Individual systems} \subsubsection{S\,1063 and S\,1113} The two stars below the subgiant branch, S\,1063 and S\,1113, both show relatively strong \ion{Ca}{ii} H\&K emission, and are the only two stars in our sample showing H$\alpha$ in emission, shown in Fig.~\ref{subha}. We use the orbital solutions for both objects to try and identify the star responsible for these emission lines. The velocities of both components in S\,1113, and of one component in S\,1063 are indicated in Fig.~\ref{subha}. The H$\alpha$ line profile of S\,1063 is asymmetric, showing emission which is blue-shifted with respect to the absorption. The location of the absorption line is compatible with the velocity of the primary, which dominates the flux; the emission is probably due to the secondary. Remarkably, the \ion{Ca}{ii} H\&K emission peak is at the velocity of the primary. This suggests that the H$\alpha$ emission is not chromospheric in nature. The H$\alpha$ emission of S\,1063 does not show the double peak that is known to indicate accretion disk emission (Horne \&\ Marsh, 1986). \nocite{hornmars} In S\,1113 the H$\alpha$ emission profile is symmetric and broad, with full width at continuum level of 15 \AA. The emission peak is centered on the more massive star, which contributes 82\%\ of the total flux (Table~\ref{fluxes}). This suggests that the H$\alpha$ emission is due to the primary. The \ion{Ca}{ii} H\&K emission shows marginal evidence for a double peak, suggesting that both stars contribute to the chromospheric emission. In Fig.~\ref{subca} we indicate the expected position of the H\&K lines for both stars. For the phase observed, their peaks overlap in the crosscorrelation function. In the figure we use the velocities resulting from fitting the order that gives the 'cleanest' crosscorrelation. \begin{figure} \centerline{\psfig{figure=subha.ps,width=\columnwidth,clip=t}} \caption{H$\alpha$ in S\,1113 (top), S\,1063 (middle) and comparison giant S\,1288. The intensity is normalized to the continuum level; the upper two spectra are displaced vertically by 0.5 and 1 unit. The line marks the radial velocity shift of the stars as determined by the crosscorrelation; for S\,1113 the primary (P) and secondary (S) star are separately indicated.} \label{subha} \end{figure} \begin{figure} \centerline{\psfig{figure=s1113ca.ps,width=\columnwidth,angle=-90,clip=t}} \caption{\ion{Ca}{ii} H\&K emission cores in S\,1113. The vertical lines mark the shifted position of the lines for the primary (P) and secondary (S) as derived from the crosscorrelation with the radial velocity standard \object{HD$\,$132737} (K0III, ESA 1997). Errors are also indicated. The long horizontal lines indicate the upper and lower limit chosen to determine the emitted \ion{Ca}{ii} H\&K flux (see Sect.~3.1).} \label{subca} \end{figure} \subsubsection{S\,1072 and S\,1237} The \ion{Ca}{ii} H\&K emission in the wide binaries S\,1072 and S\,1237 is only marginally significant. The level of their X-ray and \ion{Ca}{ii} emission is more appropriate for active main-sequence stars (Fig.\ref{xca}). One might speculate that it is due to the invisible companion of the giant detected in the crosscorrelation; even if this were the case, we would not understand why this companion would be chromospherically active. We conclude that we do not understand why these two stars are X-ray sources. We find no indication for a faint secondary in the crosscorrelation profile of S\,1237; at the time of observation, the spectra of two equally massive stars as suggested by Janes \&\ Smith (1984) would be separated by 4 to 7 km s$^{-1}$ (derived from the ephemeris in Mathieu et al. 1990). Since the secondary is 1.6 magnitude fainter in V than the primary, we think that this small separation is compatible with finding a single peak in the crosscorrelation. \subsubsection{S\,1242} S\,1242 is chromospherically active, as shown by its \ion{Ca}{ii} H\&K emission. We suggest that this activity, which also explains the X-rays, is due to rapid rotation induced by tidal interaction at periastron, which tries to bring the subgiant into corotation with the orbit at periastron. If we assume that the observed period of photometric variability is the rotation period we derive an inclination of $\sim$ 9$\degr$ using our maximum value of $v\sin i$ and the estimated radius. This would be in agreement with a companion at the high end of the range 0.14--0.94 $M_{\odot}$ allowed by the mass function (Mathieu et al. 1990). \subsubsection{S\,1040} Our detection of clear chromospheric emission indicates that the X-ray emission of S\,1040 is due to the giant. The white dwarf has a low temperature, and is unlikely to contribute to the X-ray flux. We find a rather slow rotational velocity for the giant, 3\,km s$^{-1}$. Gilliland et al.\ (1991) detected a periodicity of 7.97 days in the visual flux (B and V bandpasses) of S\,1040, with an amplitude of 0.012\,mag. If this is the rotation period of the giant, the radius of 5.1\,$R_{\sun}$ (Landsman et al.\ 1997) implies an equatorial rotation velocity of $v=32$\,km s$^{-1}$. This is compatible with the velocity measured with our crosscorrelation, $v\sin i$, for an inclination $i\ltap5.3\degr$. This inclination has an a priori probability less than 0.5\%; and it implies an unacceptably high mass for the white dwarf, from the measured mass function $f(m)=0.00268$. We conclude that the 8\,days period cannot be the rotation period of the giant. It is doubtful that the white dwarf can be responsible, as its contribution to the $B$ and $V$ flux is small. \subsubsection{S\,1082} The H$\alpha$ absorption profile of the blue straggler S\,1082 is variable. If we consider the most symmetric spectrum profile, that of 00:01 UT, as the unperturbed profile of the primary, we find that the changes are due to extra emission. This is illustrated in Fig.~\ref{s1082ha}. We suggest that this variation is due to the subluminous companion, possibly to a wind of that star. We have also investigated the presence of a broad shallow depression underlying the \ion{Na}{i} D lines (near $\lambda$ 5895 \AA) and the \ion{O}{i} triplet (near $\lambda$ 7775 \AA) as found by Mathys (1991). We find that this broad component is variable, as illustrated in Fig.~\ref{s1082na}. Mathys (1991) suggests that the broad component originates in the subluminous companion. This companion outshines the primary by a factor six at $\lambda$ 1500 \AA\, and thus is presumably hot (Landsman et al. 1998). We note that the star cannot be too hot or it would not show neutral lines. \nocite{landea98} \begin{figure} \centerline{\psfig{figure=haflux_s1082.ps,angle=-90,width=\columnwidth,clip=t}} \caption{H$\alpha$ profile for the three exposures of S\,1082. The spectra are labelled according to the UT at start of the exposure. For clarity, the lower spectrum is offset with $-0.2 \times 10^{-13}$ erg s$^{-1}$ \AA$^{-1}$ cm$^{-2}$ in both figures. The lower panels show the difference between the first and second (left) and the first and the third exposure.} \label{s1082ha} \end{figure} \begin{figure} \centerline{\psfig{figure=s1082na.ps,angle=-90,width=\columnwidth,clip=t}} \caption{\ion{Na}{i} D lines in S\,1082. Shown from left to right are the three spectra labelled according to the UT at start of the exposure. Variability is most clearly seen left of the Na $\lambda$ 5895.92 line.} \label{s1082na} \end{figure} \section{Discussion and conclusions} In this paper we have tried to find an explanation for the X-ray emission of seven sources in \object{M$\,$67}. For S\,1242 and S\,1040 we have concluded from the \ion{Ca}{ii} H\&K emission cores that magnetic activity is responsible for the X-rays. This is supported by filling in of H$\alpha$ (see e.g. Montes et al. 1997; Eker et al. 1995). In S\,1242, activity is likely to be triggered by interaction at periastron in the eccentric orbit. This is also reflected in the period of photometric variability. For S\,1040, the reason for activity is less clear. The explanation could involve mass transfer from the precursor of the white dwarf to the giant and the latter's subsequent expansion during the giant phase. As was already noted by Landsman et al. (1997), a similar system is \object{AY Cet}, a binary of a white dwarf of $T_\mathrm{eff}=18\,000$ K (Simon et al. 1985) and a G5III giant in an 56.8 days circular orbit. The $v\sin i$ of that giant is also low, 4 km s$^{-1}$, and the long photometric period of 77.2 days implies asynchronous rotation (Strassmeier et al. 1993). The X-ray luminosity for \object{AY Cet} is $1.5 \times 10^{31}$ erg s$^{-1}$ in the 0.2--4 keV band as measured with Einstein by Walter \&\ Bowyer (1981), somewhat higher than the luminosity of S\,1040. (With the coronal model discussed by Belloni et al. (1998), the countrate for S\,1040 corresponds to $5.6 \times 10^{30}$ erg s$^{-1}$ in the 0.2--4 keV band.) Walter \&\ Bowyer attribute the X-rays to coronal activity of the giant. \nocite{simoea} \nocite{waltbowy} \nocite{ekerea} \nocite{mfea} The \ion{Ca}{ii} H\&K emission cores in S\,1063 and S\,1113 are very strong. In S\,1113 we might even see emission of both stars. Due to the shape of the H$\alpha$ emission we cannot conclude with certainty that the X-rays arise in an active corona and not in a disk or stream. The wings in the emission peak of S\,1113 are very broad. However, Montes et al. (1997) have demonstrated that the excess emission in the H$\alpha$ lines of the more active binaries is sometimes a composite of a narrow and broad component, the latter having a full width at half maximum of up to 470 km s$^{-1}$. They ascribe this broad component to microflaring accompanied by large scale motions. We note the similarity between S\,1113 and \object{V711 Tau}, a well known extremely active binary of a G5IV ($v\sin i$ = 13 km s$^{-1}$) and K1IV ($v\sin i = 38$ km s$^{-1}$) star in a 2.84 days circular orbit and a mass ratio 0.79 (Strassmeier et al. 1993); the mass ratio of S\,1113 is 0.70 (Mathieu et al. 1998, in preparation). From the countrate of \object{V711 Tau} in Table~\ref{tabros} we find $L_{\mathrm{x}}=6.8 \times 10^{30}$ erg s$^{-1}$ in the 0.1--2.4 keV band using the same model as in Belloni et al. (1998) with $N_{H}=0$, which is comparable to the luminosity of S\,1113 in the same band $L_{\mathrm{x}}=7.3 \times 10^{30}$ erg s$^{-1}$ (Belloni et al. 1998). The H$\alpha$ emission of S\,1063 is more difficult to explain. As this system is not double-lined, H$\alpha$ emisson by the (invisible) secondary star would have to be strong to rise above the continuum of the primary. \nocite{mfea} In the binaries S\,1072 and S\,1237 we see no H$\alpha$ emission while the level of \ion{Ca}{ii} H\&K emission is low in comparison with active stars of the same luminosity class. We have no explanation for this. For S\,1072, an option is a wrong identification of the X-ray source with an optical counterpart. Belloni et al. (1998) give a probability of 43\% that one or two of their twelve identifications of an X-ray source with a binary in \object{M$\,$67} is due to chance. No \ion{Ca}{ii} H\&K emission is seen in the spectrum of the blue straggler S\,1082. Possibly, the X-ray emission has to do with the hot, subluminous secondary that could also cause the photometric variability and whose signature we might have seen in the H$\alpha$ line. \begin{acknowledgements} The authors wish to thank G. Geertsema for her help during our observations, P. Groot for providing his program to compute projected rotational velocities with the Fourier-Bessel transformation method and M. van Kerkwijk for comments on the manuscript. MvdB is supported by the Netherlands Organization for Scientific Research (NWO). \end{acknowledgements}
\section{Introduction} Since its construction the protons and anti-protons used in the Tevatron Collider have been produced and injected into the Tevatron, by the Main Ring which was the original 400 GeV proton synchrotron constructed during the 1970's. The Main Injector is a new 120-150 GeV rapid cycling synchrotron in a tunnel separate from that of the Tevatron. The Main Injector improves significantly the anti-proton production capability of the complex. The new 8 GeV permanent magnet storage ring, the Recycler, in the Main Injector tunnel, permits reuse of the anti-protons that remain in the Tevatron at the end of a store. Since most of the luminosity degradation results from beam dilution rather than anti-proton loss, this effectively doubles the available anti-protons for collisions. The greatly increased luminosity of the Tevatron Collider opens up windows on new physics. The production rate for heavy objects such as the top quark will be greatly enhanced. Perhaps most strikingly the increased luminosity may eventually provide sensitivity to the Higgs boson system which is thought to generate the mass of the observed particles, the $W$ boson, the $Z$ boson, and the quarks and leptons. The Main Injector is constructed with extracted beams that can be used for fixed target experiments. The operational possibilities are sufficiently flexible to interleave anti-proton production and fixed target physics with modest impact on either. One of the primary uses of the high intensity extracted proton beam will be to produce a beam of neutrinos which will then be directed towards the Soudan mine in Minnesota where a distant detector can make measurements on the evolution of the neutrino species in the beam. The couplings of the three generations of quarks are described using the Cabibbo-Kobayashi-Maskawa flavor mixing matrix. A complete understanding this matrix does not yet exist. For example, the violation of Charge-Parity symmetry has been observed in the decay of neutral long-lived $K^{0}_{L}$ mesons but nowhere else. Kaon beams from the Main Injector promise key measurements in that system. The analogous investigation of the $b$-quark system is only beginning. The collider experiments CDF and D\O\ or/and a new dedicated experiment, BTeV, have ready access to large numbers of B hadrons, thus permitting extensive measurements beyond those being made at electron-positron B factories. The varieties of physics in these different sectors are all at or beyond the edge of our theoretical understanding and offer new insight on our world. In the rest of this paper we explore the machine and the enabled physics in a little more detail. \section{The Machine} The Main Injector is a rapid cycling proton synchrotron with a circumference half that of the Tevatron. Protons are injected from the Booster machine to the Main Injector at 8 GeV and are accelerated to 120 GeV for either extraction to the anti-proton production target or to an external physics target. For injection of either protons or anti-protons into the Tevatron, the Main Injector ramps to 150 GeV. The machine performance parameters of interest for the particle physicist are given in Table~\ref{table-miparam}. In mixed modes of operation, the Main Injector can deliver $2.5$~$10^{13}$ protons to the experimental target and $5.0$~$10^{12}$ protons to the anti-proton production target every 2.87 secs. This causes a 15-20\% reduction of anti-proton production as a result of the increased cycle time. There are also potential improvements which could ultimately yield $5-10$~$10^{13}$ protons per cycle. \begin{table} \caption{Main Injector Performance Characteristics.} \begin{tabular}{l|ccc} &$\overline{p}$ Production &Fast Spill &Slow Spill\\ \tableline Energy(GeV) &120 &120 &120\\ Protons per Cycle & $5.0$~$10^{12}$ & $3.0$~$10^{13}$ & $3.0$~$10^{13}$ \\ Flat Top(secs)& 0.01 &0.01 &1.00\\ Cycle Time(secs)& 1.47 & 1.87 & 2.87 \ \label{table-miparam} \end{tabular} \end{table} The Collider luminosity is controlled by the total number of anti-protons available to accelerate and store. The anti-protons are produced by the 120 GeV protons incident on a nickel target. The produced anti-protons are focussed with a Lithium lens and collected in a debuncher ring at 8 GeV. From the debuncher they are transferred to the accumulator ring, also at 8 GeV, where they are cooled, thus producing stored anti-proton bunches. These are transferred to the 8 GeV Recycler ring before acceleration in the Main Injector, and eventually the Tevatron, to the full energy. During a store, nuclear collisions cause some attrition while the beam-beam effects lead to an increase in the emittance of the beams. This effect dominates the reduction in luminosity over a period of hours. The effect is most strong on the anti-protons since the proton bunches are, in general, more intense than those of anti-protons. At the end of a store, which is usually defined by the luminosity having dropped to a few tenths of its initial value, the number of anti-protons has only reduced by a factor of two or less. These survivors can be decelerated through the Tevatron and Main Injector and captured in the Recycler. This is an 8 GeV permanent magnet machine equipped with stochastic cooling. It permits the recovery of emittances suitable for reinjection into the Tevatron and thus results in an effective factor of two enhancement to the number of available anti-protons under normal operation. The Tevatron Collider will operate with the luminosities indicated in Table~\ref{table-tevchar} \begin{table} \caption{Tevatron Collider Operating Characteristics.} \begin{tabular}{l|cccc} &Bunch Spacing(nsec) & Inst. Luminosity($\rm{10^{31} cm^{2} sec^{-1}}$) &Interactions per crossing&Luminous Region(cm)\\ \tableline Run Ib (1994-6) & 3500 & 1.6\tablenote{This was typical, the absolute record exceeded 2.5 $\rm{10^{31} cm^{2} sec^{-1}}$.} & 1-2 & 30 \\ Run II (2000-3) & 396/132 & 10/20 & 1-2/1-2& 30/15 \\ Run III(2004-7) & 132 & 50 & 5 & 15 \ \label{table-tevchar} \end{tabular} \end{table} Further enhancements can be expected from the introduction of electron cooling in the Recycler and from the use of tune compensation in the Tevatron. A possible accumulation of integrated luminosity as a function of time is shown in Table~\ref{table-integL}. Before the full operation of the Large Hadron Collider at CERN, more than 10~$\rm{fb^{-1}}$ could be expected. \begin{table} \caption{Tevatron Collider Integrated Luminosity.} \begin{tabular}{l|cccccccc} Year &2000&2001&2002&2003&2004&2005&2006&2007\\ \tableline Peak Luminosity($10^{31} cm^{2} sec^{-1}$)& 5 & 10 &20 && 40 & 50 & 50 & 50 \\ Integrated Luminosity($\rm{fb^{-1}}$)&0.5&1.0&2.0&&4.5&5.5&5.5&5.5\\ Accumulated Luminosity($\rm{fb^{-1}}$)&0.5&1.5&3.5&&8.0&13.5&19.0&24.0\ \label{table-integL} \end{tabular} \end{table} The Main Injector was being commissioned\cite{bd-opspage} at the time of the conference and that process has gone extremely well. In particular the machine is being operated with parameter values near those of the design and with relatively minimal use of correction elements. As for performance, it is already close to its intensity and cycle time goals. Meanwhile the Recycler is still in the final stages of installation. Beam has been passed through a fraction of its circumference. \section{The Physics} \subsection{Neutrino Mixing and mass} Over the years there have been speculations about whether or not neutrinos have identically zero mass. If not, one expects the weak eigenstates states to mix so that a pure beam of neutrinos of one species will evolve to contain an admixture of neutrinos of one or more other species. This phenomenon is called neutrino oscillation. The probability that a transition takes place is proportional to $\sin^{2}(1.27{\Delta{m}^2}{L}/E)$ where the difference in mass squared between the two neutrino states, $ {\Delta{m}^2}$, is measured in $(eV)^{2}$, the path length, L, in kilometres and the energy, E, in GeV. The strength of the oscillations is usually described by a factor $\sin^{2}{2\theta}$. At the present time there are a number of observations\cite{feldman} from experiments which could be explained had the the neutrinos a finite mass. However the picture is quite complicated. The observation of a deficit of neutrinos from the sun suggests oscillations with very low $ {\Delta{m}^2}$. The experiments measuring the fluxes of atmospheric neutrinos, including the recent results from Kamiokande\cite{superkhere}, suggest ${\Delta{m}^2}\simeq 10^{-3} - 10^{-2}(eV)^{2}$ with maximal strength. The LSND experiment at Los Alamos has observed a hint of oscillations with $ {\Delta{m}^2}\simeq 10^{-2} - 10^{0}(eV)^{2}$ with $\sin^{2}{2\theta}$ as low as $10^{-3}$. These observations are indicated in Fig.~\ref{fig-allhints}\cite{conrad_ichep}. \begin{figure}[ht] \vskip -0.5 cm \centerline{\epsfxsize 2.8 truein \epsfbox{montallhints_bare.ps}} \vskip -1.0 cm \caption[]{ \label{fig-allhints} \small Hints of neutrino oscillations from present measurements. Note that the most recent atmospheric measurements, reported at this conference, suggest a region of ${\Delta{m}^2}$ somehwat higher than indicated in this figure from the summer of 1998. } \end{figure} The NuMI project, Neutrinos at the Main Injector, will construct a neutrino beam with energies peaking in the 10-25 GeV range. These neutrinos will be directed at two detectors, one on the Fermilab site, the other 740 km further north in the Soudan mine in Minnesota, see Fig.~\ref{fig-numimap}. \begin{figure}[ht] \centerline{\epsfxsize 1.9 truein \epsfbox{montminosmap.eps}} \vskip 0 cm \caption[]{ \label{fig-numimap} \small Map indicating the trajectory of the neutrino beam from Fermilab to the Soudan mine in Minnesota. Inset at the bottom is a ``section'' of the earth showing the penetration of the neutrino trajectories through the earth's crust.} \end{figure} The MINOS experiment\cite{minos_expt}, Main Injector Neutrino Oscillation, will comprise two detectors one on the Fermilab site which monitors the neutrino beam interactions close to the source and one at the Soudan mine. The far detector sketched in Fig.~\ref{fig-minosfardet} will consist of iron toroid plates and solid scintillator sheets. The goals of the experiment are unequivocally to observe the oscillations indicated by the SuperKamiokande and other experiments and to identify the mode(s) of oscillation. Neutral current events are distinguished from charged current events using measurements of the shapes of the neutrino induced showers. This technique gives some promise of positive identification of oscillations into $\nu_{\tau}$. This capability could be enhanced by a supplementary emulsion detector should the existing measurements continue to suggest that the $\nu_{\tau}$ mode is the relevant one. The sensitivity of the experiment is primarily at high values of $\sin^{2}{2\theta}$ and with $ {\Delta{m}^2}\ge 10^{-3}(eV)^{2}$. \begin{figure}[ht] \vskip -11.5 cm \centerline{\epsfxsize 3.3 truein \epsfbox{montfardet_rot.ps }} \vskip 9.5 cm \caption[]{ \label{fig-minosfardet} \small Layout of the MINOS ``Far Detector'' which will be situated in the Soudan mine.} \end{figure} The mini-BooNE experiment\cite{boone}, Booster Neutrino Experiment, is not a Main Injector experiment. Rather it uses protons from the 8 GeV proton Booster machine to generate low energy neutrinos. Using an apparatus derived from that of LSND and situated about 1 km from the source, it aims definitively to cover that region of parameter space corresponding to the LSND observations. The systematic uncertainties would be significantly different from those of LSND. \subsection{Physics of the Kaon System} The Cabbibo-Kobayashi-Maskawa(CKM) matrix has nine elements that can be described using four real parameters of which one is a phase angle. In turn, if the matrix is unitary, these parameters can be represented by a triangles. The lengths of the sides are controlled by the various transition amplitudes. The magnitude of Charge-Parity(CP) symmetry violation is controlled by the phase angle. This description of the quarks and their couplings may or may not hold in nature. It is one of the highest priorities of high energy physics to explore the CKM matrix in more detail and to determine whether or not the conjectures about its properties are true. At present the single indication of CP symmetry violation, which is what requires the complex matrix element, occurs in the $K^{0}_{L}$ system\cite{blucher}. At present CP violation has not been observed in any other system containing strange quarks nor yet definitively observed in the $b$-quark sector\cite{cdf-sin2beta}. The conditions for existence of CP violation in any given system may be quite involved. In the $K^{0}_{L}$ system, for example, at the time of the conference it was still possible for the observed CP violation to be completely described by the mixing effects which are controlled by the parameter $\epsilon$. The search is for CP violation in the decay, `` direct CP violation'', which is controlled by the parameter $\epsilon$'. The new measurement\cite{blucher,ktev_new}, which appeared since the time of the conference, give $\cal{R}e(\epsilon$'/$\epsilon)$$\simeq (28 \pm 4)\times10^{-4}$ \begin{figure}[ht] \centerline{\epsfxsize 5.5 truein \epsfbox{montburas_triangle.ps}} \vskip -.2 cm \caption[]{ \label{fig-burasfig} \small The Unitarity Triangle associated with the Cabbibo-Kobayashi-Maskawa flavor mixing matrix and the measurements possible in the kaon system.} \end{figure} In order to make progress, thoughts have turned to other possibilities\cite{cooper-kaons}. A version\cite{buras} of the unitarity triangle is shown in Fig.~\ref{fig-burasfig}. In principle it is possible to over-constrain the triangle and hence to test the theory using only measurements with kaons. As indicated, a measurement of the branching fraction for $K^{0}_{L}\rightarrow\pi^{0}\nu\overline{\nu}$ would directly constrain the height of the triangle while a similar measurement of the charged decay $K^{+}\rightarrow\pi^{+}\nu\overline{\nu}$ determines the radius of an arc which should also pass through the apex of the triangle, if the theory is correct. The charged kaon decay has been sought\cite{bnl-kaplus} at Brookhaven national Laboratory in the decays of stopped kaons with one event observed. Recently a proposal\cite{fnal-kaplus}, the ``CKM'' experiment, has been made to use decays in flight. The apparatus is shown in Fig.~\ref{fig-ckmexp}; the beam would be a 22 GeV radiofrequency-separated beam of charged kaons at the Main Injector. It is interesting to look carefully at the aspect ratio of the experiment. It is very long and very narrow approximating an instrumented sewer pipe. The goal is to fully identify and measure the incident kaon and the outgoing charged pion, the only two measurable particles in the process. \begin{figure}[ht] \centerline{\epsfxsize 4.4 truein \epsfbox{montckm_apparatus.ps}} \vskip 0 cm \caption[]{ \label{fig-ckmexp} \small Layout of the CKM Experiment.} \end{figure} The equivalent neutral kaon experiment\cite{fnal-kami}, ``KaMI'', which would search for $K^{0}_{L}\rightarrow\pi^{0}\nu\overline{\nu}$ could be derived from the KTeV experiment making the recent measurements of $\epsilon$'/$\epsilon$ at Fermilab. The key elements are the electromagnetic calorimeter and the photon vetos which are crucial to the background suppression. As with the charged kaon decay, this would be a very difficult measurement. \begin{figure}[ht] \centerline{\epsfxsize 4.4 truein \epsfbox{montkami_app.ps}} \vskip 0 cm \caption[]{ \label{fig-kamiexp} \small Layout of the KAMI Experiment} \end{figure} Completing the suite of kaon proposals is the ``CPT'' Experiment\cite{fnal-cpt}. The goal is to measure CP violation in a number of modes especially in $K^{0}_{S}$ decays. It also gives the opportunity to measure the phase of the charged pion decays which in conjunction with the measurement of $\epsilon$'/$\epsilon$ gives a check on CPT symmetry with a sensitivity which corresponds to the Planck scale. It should be noted, see Fig.~\ref{fig-cptexp}, that since it is necessary to measure the interference terms between $K^{0}_{L}$ and $K^{0}_{S}$, the apparatus would be significantly shorter than either of the other two experiments. The $K^{0}$ beam is derived from the same RF-separated $K^{+}$ beam that is used for the ``CKM'' experiment. \begin{figure}[ht] \centerline{\epsfxsize 2.5 truein \epsfbox{montcpt_detector.eps}} \vskip 0 cm \caption[]{ \label{fig-cptexp} \small Layout of the CPT Experiment.} \end{figure} \samepage \subsection{Physics of the B System} In order to access large numbers of $b$ quarks at Fermilab, it is necessary to turn to the Tevatron Collider. The existing two detectors, CDF, see Fig.~\ref{fig-cdf}, and D\O\, see Fig.~\ref{fig-d0}, are being upgraded for operation in the Main Injector era. The upgrades\cite{cdf-upgrade,d0-upgrade} are extensive and are driven by the physics goals and by the changed operating characteristics of the Tevatron; the decreased spacing between bunch crossings has forced a rework of all the front-end eleectronic systems to introduce pipelines. A particular region of improvement is in the tracking in which each detector is being substantially modified. D\O\ has installed a central solenoid and both experiments are constructing new outer trackers. The inner, silicon, systems will have upward of 600,000 channels each. These will provide high quality $b$-quark tagging and B meson reconstruction. In addition to enhancing the B-physics capabilities, $b$-quark tagging is recognised as an important tool in the exploration of and search for higher mass states. This was demonstrated in top-quark physics and is expected to be true for Higgs, SUSY or technicolor states. \begin{figure}[h] \vskip 0.2 cm \centerline{\hspace{3cm}\epsfxsize 5.5 truein \epsfbox{montcdf_side.eps}} \vskip 0.5 cm \caption[]{ \label{fig-cdf} \small The upgraded CDF detector. } \end{figure} \begin{figure}[h] \centerline{\hspace{1.5in}\epsfxsize 6.0 truein \epsfbox{montd0detect_tomd.ps}} \vskip -2 cm \caption[]{ \label{fig-d0} \small The upgraded D\O\ apparatus. } \end{figure} The physics which will be addressed by CDF and D\O\ is wide ranging. In later sections electroweak studies and physics beyond the standard model will be addressed but in this section we will concentrate on B physics. Recent results from CDF presage a bright future for the general purpose detectors at the Tevatron and for B physics at hadron colliders more generally. There is also an initiative to consider a dedicated detector\cite{btev-proposal}. The BTeV proposal is motivated by the enormous production rate for $b$ quarks and the potential of a detector which accentuates the forward and backward directions in order to exploit the favorable mapping of rapidity to solid angle and to benefit from the relatively larger decay lengths for $b$ quarks of a given transverse momentum. The planned detector is shown in Fig.~\ref{fig-btev}. It consists of two spectrometer arms. Each arm is equipped with a silicon detector system inside the beam pipe and ring imaging Cherenkov counters. It is expected that the latter will be especially important. \begin{figure}[h] \centerline{\epsfxsize 4.4 truein \epsfbox{montbtev.eps}} \vskip 0 cm \caption[]{ \label{fig-btev} \small The BTeV apparatus. } \end{figure} If the standard model is correct, the same triangle in Fig.~\ref{fig-burasfig} should describe the $b$-quark system. Measurements of the CP violating asymmetry in the decay $ B \rightarrow J/\psi K^{0}_{S}$ would determine $\sin 2\beta$. The recent measurement\cite{cdf-sin2beta} from CDF finds $\sin 2\beta=0.79_{-0.44}^{+0.41}$ suggesting a positive value at about the 90\% C.L. A feature of this measurement is the use of several different flavor tagging techniques. With approximately 2$\rm{fb^{-1}}$ and the upgraded detectors, the uncertainty on $\sin 2\beta$ will be reduced below 0.1 for each experiment. Similar uncertainties are projected for $\sin 2\alpha$ although the interpretation for this case is considered to be more difficult. Measurement of the third angle, $\gamma$, will be a challenge. During the last year, there have been a number of results from CDF on various aspects of higher mass B hadron states. The measurements of $B_s$ mixing\cite{cdf_b_s} are competitive with those from LEP and SLD. Extrapolating to Run II, we expect sensitivities in the range $x_s \ge 25$ from each experiment thus comfortably covering the expected range. The observation\cite{cdf_b_c} of the $B_c$ meson has further demonstrated that sophisticated studies of B physics are possible at a hadron collider. One should note that the final state used for this observation was semi-leptonic. Despite the incomplete reconstruction the backgrounds were manageable. Given the enormous rates, the Tevatron is arguably the best place to do B physics. This is especially true for states which are not decay products of the $\Upsilon_{4S}$ and are therefore difficult to access with an electron-positron B factory. \subsection{Electroweak Physics} The hadron colliders have for some time contributed to the suite of measurements which provide stringent cross-checks of the electroweak model. The initial measurements from the CERN $Sp\overline{p}S$ observation of the $W$ and $Z$ bosons paved the way. The current measurements\cite{wmass_measurements} from the Tevatron keep pace in precision with those from LEP and at each new level of precision we see ways to constrain the details of the production measurements from the data themselves. For example in the recent D\O\ measurement it was found that the constraint from the rapidity distribution of the bosons is only marginally weaker than that from the parton distribution function measurements from the worlds experiments. The current errors are 80-90 MeV per experiment. Already the latter are strongly influenced by measurements of the $W$ asymmetry at the Tevatron. The constraints from the data themselves will scale with statistics to higher integrated luminosity. This means that the current estimates of about 40 MeV per channel and per experiment are indeed possible. \begin{figure}[] \centerline{\epsfxsize 2.6 truein \epsfbox{montmarcel_wmass_linear.eps}} \vskip 0 cm \caption[]{ \label{fig-wmassevol} \small Expected evolution of the precision of a measurement of the $W$-boson mass at the Tevatron Collider. } \end{figure} \begin{table} \caption{Uncertainties in a single experiment top mass determination.} \begin{tabular}{l|cc} Uncertainty(GeV) & Run I & Run II\\ \tableline Statistical& 5.6 & 1.3 \\ \tableline Jet Energy Calibration & 4.0 & 0.4 \\ Gluon ISR/FSR & 3.1 & 0.7 \\ Detector Noise etc & 1.6 & 0.4 \\ Fit Procedure & 1.3 & 0.3 \\ \tableline All Systematic & 5.5 & 0.9 \\ \tableline Total & 7.8 & 1.6 \ \label{table-topmass} \end{tabular} \end{table} This kind of evolution as expressed in an earlier study\cite{tev2000} is illustrated in Fig.~\ref{fig-wmassevol}. That study considered particularly the effects of the underlying event on the technique which uses the transverse mass of the event (lepton plus neutrino) as the primary measure of the boson mass. This leads to a relative deterioration as the number of interactions per bunch crossing increases. What is also shown is the subsequent evolution as we decrease the bunch spacing in the machine to 132 nsecs. Fora given luminosity this decreases the number of interactions per crossing. We are also learning to take advantage of all the measures of the boson mass, the lepton transverse momentum, and the neutrino transverse momentum. It is gratifying that this projection made some four years ago holds good through the 100 $\rm{pb}^{-1}$ of data from which the recent measurements are derived. A very powerful newcomer to precision measurements is the top mass. It enters into the electroweak parameters through its dominance of the quark loops. From the existing data, the 3\% uncertainty makes the top-quark mass the best known of all quark masses. The 5 GeV uncertainty in the top-quark mass is equivalent to about 30 MeV uncertainty in the $W$ mass in terms of its its sensitivity to the Higgs mass.. Currently the dominant error\cite{topmasserrors} comes from the calibration of the jets. Recently CDF observed the $Z\rightarrow b\overline{b}$ decay. Extrapolating to the luminosities and upgraded detectors with silicon track triggers to enhance the sensitivity to this channel, a very precise calibration of the jet energies becomes possible. This will be used in conjunction with the $W$-boson decays to jets which are present in the top signal data themselves. The resulting evolution of a single experiment in the lepton-plus-jets channel is illustrated in Table~\ref{table-topmass}. An uncertainty of less than 2 GeV appears to be possible. A number of other electroweak studies are possible including the comparison of the $W$ width as determined by direct and indirect methods and a study of the $Z$ asymmetry. The latter may serve either as a determination of $\sin^{2}\theta_{W}$ for the light quarks or another constraint on the parton distribution functions. Taken together, the masses of the $W$ boson and the top quark lead to constraints on the mass of the Higgs boson either in the standard model or in the supersymmetric variants. This is illustrated in Fig.~\ref{fig-mwmt} where, in the plane of the $W$ mass and the top-quark mass, the various measurements including the latest from CDF and D\O\ at the Tevatron. We can anticipate that these indirect measurements will determine the Higgs mass to about 50\% of its value. \begin{figure}[h] \centerline{\epsfxsize 4.0 truein \epsfbox{montmwmt3mar99.eps}} \vskip 0 cm \caption[]{ \label{fig-mwmt} \small $M_{W}$ versus $M_{t}$. } \end{figure} \subsection{Physics beyond the Standard Model} There are as many searches for new physics as can be generated by the imagination of physicists. In the search for compositeness, structure, higher mass bosons, leptoquarks, the present limits are in the few-hundred-GeV range. With $2~\rm{fb}^{-1}$ these searches will be sensitive close to the 1 TeV range. Indirect searches as exemplified by the measurement of the Drell-Yan, lepton pair cross section and the dijet mass spectrum, are sensitive, through possible contact terms to the 5 TeV range. \begin{figure}[] \centerline{\epsfxsize 3.3 truein \epsfbox{monttechnicolor.ps} } \vskip 0 cm \caption[]{ \label{fig-technicolor} \small Mass spectra expected for technicolor signals for the $\pi_{T}$ and the $\rho_{T}$ .} \end{figure} \subsubsection{Technicolor} A longstanding candidate to explain the origin of electroweak symmetry breaking is a new strong interaction, technicolor, analogous to QCD. Such a strong interaction would lead to massive electroweak bosons in a manner analogous to the way the masses of the pion and $\rho$ meson are generated by QCD. There are complications with the simplest forms of such a theory but variations\cite{technicolorsummary} of the original scheme continue to be explored. \begin{table}[h] \caption{Mass ranges covered for a $5\sigma$ discovery in SUSY models.} \begin{tabular}{l|c|cc} Model & SUSY Particle & Run I( 0.1 $fb^{-1}$ & Run II(2.0 $fb^{-1}$\\ & & Mass Limit(GeV) & Mass Limit(GeV) \\ \tableline SUGRA &&\\ \tableline & $\tilde{\chi}_{1}^{\pm}$ & 70\tablenote{95\% CL.} & 210 \\ & $\tilde{g}$ & 270\tablenote{95\% CL.} & 390 \\ & $\tilde{t}_{1}(\rightarrow b\tilde{\chi}_{1}^{\pm})$& & 170 \\ \tableline GMSB&&\\ \tableline & $\tilde{\chi}_{1}^{\pm}$ & & 265 \\ & $\tilde{\tau}$ & & 120 \ \label{table-susy} \end{tabular} \end{table} The production cross section for the some of the states can be quite substantial\cite{womestilane}. Analysis of Monte Carlo simulations of technicolor signals using a basic signature of a $W$, seen through its leptonic decay, along with two jets gives a dijet mass spectrum as shown in the upper plot of Fig.~\ref{fig-technicolor}. Some topological requirements are then applied and a $b$-quark tag is requested. This leads to the middle of the three plots. The dijet spectrum clearly shows an excess which corresponds to the technipion decaying to two $b$-quark jets. In the bottom plot, the mass of the combined $W$ boson and the two-jet system shows a peak over background corresponding to the technirho. Once again the importance of $b$-quark tagging techniques is demonstrated. \subsubsection{SUSY} The mainstream of theoretical thinking with respect to the physics of electroweak symmetry breaking and the physics above 100 GeV is dominated by those who consider that supersymmetry, SUSY, should play a strong role. SUSY comes in may guises but always leads to a proliferation of postulated particles differing in spin by one half unit with respect to the ``standard particles''. Thus there is a spin-one-half gluino, the partner of the gluon, a spin-one-half photino, the partner of the photon and a host of squarks and sleptons. The simplest assumption is that these sparticles can only be produced in pairs, a restriction that is usually formulated as conservation of a quantum number R-parity. Given R-parity conservation there is always a stable lightest supersymmetric partner(LSP) which is neutral in most theories. This leads to the presence of missing transverse energy as the most generic of SUSY indicators. As with technicolor, the use of $b$-quark tagging also can be a useful discriminator against background. Since we have not seen supersymmetric partners with the same masses as the ordinary particles, SUSY, if it exists must be a broken symmetry. The mechanism by which it is broken at very high mass scales distinguishes different models. Very commonly considered is the super gravity(SUGRA) class. In the last few years, alternatives such as gauge mediated models(GMSB) have been prominent. The latter are characterised by the gravitino($\tilde{G}$) being the LSP and cascades of decays such as $\chi_{1}^{0}\rightarrow \gamma\tilde{G}$ which generate final states with one or more photons and missing transverse energy. As indicated in Table~\ref{table-susy}, the present limits for different sparticles range up to a couple of hundred GeV for some but around 100 GeV for others. The afficionados of SUSY tend to expect that it would appear with sparticle masses in the range below 1 TeV if it is to be relevant to the electroweak symmetry breaking problem. The recent studies\cite{susy_summary}, also summarised in Table~\ref{table-susy} suggest that with 2 $\rm{fb^{-1}}$ of integrated luminosity, these ranges can be considerably extended to cover a large fraction of the ``interesting'' region. \subsection{ The Higgs Boson} Without looking beyond the physics of the standard model, it is necessary to postulate some mechanism to break the electroweak symmetry and to give mass to the $W$ and $Z$ bosons. The simplest thing to do is to assume a single complex Higgs field which in turn leads to a single neutral Higgs boson. As we have seen earlier the mass of such an object is predicted through the radiative corrections to the electroweak parameters and with the mass of the top quark and that of the W boson measured we should find the Higgs boson at the appropriate place. The search for the standard-model Higgs boson is the most widely used benchmark for the potential of planned collider experiments. Recent studies\cite{susyrun2} have put the estimates for the Tevatron Collider experiments, D\O\ and CDF, on a more solid footing. While gluon-gluon fusion has the highest Higgs production cross section, the associated , $WH$ and $ZH$, production channels offer distinctive experimental signatures through the leptonic decays of the bosons and have received much attention. The $b\overline{b}$ decay of the Higgs also adds powerful discrimination especially at low masses, below about 130 GeV, where that decay dominates. A more thorough study of the channels involving $b$-quark tagging was conducted. Further, the use of the $WW$ decay modes in conjunction with the dominant gluon-gluon fusion has been reconsidered\cite{turcothan}. The branching fractions for the latter rise strongly, see Fig.~\ref{fig-higgsbr}, with increasing Higgs mass and also are distinctive experimentally. The combined sensitivity for all channels and two experiments is shown in Fig.\ref{fig-higgsensitivity}. The figure shows the required luminosity to obtains a signal at different levels of significance, $5~\sigma$, $3~\sigma$, and the 95\% exclusion limit, as a function of Higgs mass. We see that with both experiments and 30 $fb^{-1}$ of luminosity for each experiment the sensitivity extends up to Higgs massses of 190 GeV. If the Higgs does not exist in this mass region, with 10 $fb^{-1}$ this whole region could be excluded experimentally at the 90\% C.L. \begin{figure}[] \vspace{-0.5cm} \centerline{\epsfxsize 2.43 truein \epsfbox{montsm_higgs_xsec.ps} \epsfxsize 2.2 truein \epsfbox{monthbr_ffbar.ps}\epsfxsize 2.2 truein \epsfbox{monthbr_bosons.ps} } \vskip 0 cm \caption[]{ \label{fig-higgsbr} \small Higgs-boson production cross sections and branching fractions for fermions and bosons as a function of Higgs-boson mass.} \end{figure} Other studies concentrated on the extensions of the Higgs sector, in particular the SUSY-Higgs two-doublet model. There are three neutral and two charged Higgs states and, depending on the value of the ratio of the vacuum expectation values of the two doublets, very strong coupling of the Higgs to the $b$-quark may be expected. Again the r$\rm{\hat{o}}$le of the $b$-quark tagging is important and similar sensitivities to the SUSY Higgs are achieved as in the standard-model-Higgs case. As we discussed earlier, an integrated luminosity of 20-30 $fb^{-1}$ before the startup of the LHC is anticipated. There is a considerable challenge for the experiments but even the fierce conditions may be mitigated by operating the machine in such a way as to maintain the instantaneous luminosity at or less than $5. \times {10^{32}} cm^{2} sec^{-1} $. It is clear that maximizing the exploitation of the Tevatron Collider to search for the Higgs and other new high-mass physics should be one of the highest priorities of the U.S. high energy physics program. \begin{figure}[] \centerline{\epsfxsize 3.0 truein \epsfbox{montmh_cmbned_prel.ps}} \vskip 0 cm \caption[]{ \label{fig-higgsensitivity} \small Luminosity required as a function of Higgs mass to achieve different levels of sensitivity to the standard-model Higgs boson.From the upper curve corresponds to a $5~\sigma$ discovery, the middle a $3~\sigma$ signal and the lower a 95\% exclusion limit. These limits require two experiments, Bayesian statistics are used to combine the channels and include the improved sensitivity which would come from multivariate analysis techniques.} \end{figure} \section{The Experimental Program} We have seen in the above discussion that the physics reach of the accelerator complex which we label with its newest component, The Fermilab Main Injector, is phenomenal and diverse. Full exploitation of every aspect could swallow resources in excess of what appears to be available. The attempts to construct a realistic program has so far left the various components in states of varying certitude. While some pieces are well on their way to completion of construction others remain glints in the eyes of the proponents. In order to give a sense of perspective, I have chosen to give the briefest of summaries of the status of each of the components of the Main Injector physics program. \begin{itemize} \item The Main Injector: the commissioning of the machine is well advanced; the ancillary Recycler storage ring will be completed in the coming months. \item The Tevatron Collider and the CDF and D\O\ Detectors: the Tevatron will operate in fixed target mode during 1999 and will convert to collider operations for early 2000. The upgrades of the two detectors CDF and D\O\ are in the middle of construction and completion and roll-in is expected in 2000. \item The NuMI Project: the project has approval from the appropriate authorities and a baseline for the scope, cost, funding and schedule has been approved by a Department of Energy review with funds for civil construction allocated. \item The mini-BooNe Experiment: this initial phase of a potentially longer program is an approved experiment expected to run in 2002. \item The Kaon-CP Violation experiments: these experiments, labelled, CKM, CPT and KaMI, have submitted proposals or letters of intent to the laboratory; research and development projects associated with different aspects of them have been established. \item The dedicated collider-B Experiment, BTeV: a letter of intent has been submitted and an experimental hall has been constructed; a research and development program is under way. \item A 120 GeV QCD Program: a number of groups have submitted proposals\cite{meson120} in response to the potential offered by extracted hadron beams from the the Main Injector. The primary thrust of such a program would be to emphasize QCD studies. Thus far there is no action on these proposals. \end{itemize} \section{Conclusions} The Main Injector enables a phenomenally broad and imposing array of physics and we, the field, must be wise in choosing which pieces to emphasize. Many physicists are determined to exploit the potential of this program. I am very excited to be among those physicists. \section{Acknowledgements} The talk and this paper could not have been produced with out the help many. Included among those people are Franco Bedeschi, Ed Blucher, Amber Boehnlein, Greg Bock, Janet Conrad, Peter Cooper, Marcel Demarteau, Gene Fisk, Al Goshaw, Paul Grannis, Steve Holmes, Zoltan Ligeti, John Marriner, Shekhar Mishra, Meenakshi Narain, Adam Para, Ron Ray, Maria Roco, Ken Stanfield, Gordon Thomson, Andre Turcot, Harry Weerts, Bruce Winstein, Stan Wojcicki and John Womersley who were kind enough to provide input and/or to read a draft version of the paper. To them should go the credit for the content, to me the blame for errors. This work was supported by the U.S. Department of Energy under Contract No. DE-AC02-76CHO3000.
\section{Introduction} Since faint nebulosity around quasars was discovered (Matthews \& Sandage 1963; Sandage \& Miller 1966), morphological studies of QSO host galaxies have revealed the evolutionary link between the formation of QSO hosts and the activation of QSO nucleus (Hutchings et al. 1982; Malkan 1984; Margon, Downes, \& Chanan 1984; Smith et al. 1986; Heckman et al. 1991). Photometric and spectroscopic studies of QSO hosts furthermore provided valuable clues to the nature of stellar populations of QSO host (MacKenty \& Stockton 1984; Boronson, Perrson, \& Oke 1985; Stockton \& Rigeway 1991; Dunlop et al. 1993; McLeod \& Rieke 1994). One of the most remarkable observational evidences is that galaxy interaction and merging can trigger the nuclear activities of QSOs (Stockton 1982; Hutchings \& Campbell 1983; Stockton \& Mackenty1983; Hutchings \& Neff 1992; Bahcall et al. 1997). In particular, the recent high-resolution morphological studies of QSO host galaxies by the Hubble Space Telescope (HST) and large grand-based ones found that a sizable fraction of QSO hosts have close companion galaxies likely to be interacting or merging with the hosts (Bahcall et al. 1995; Disney et al 1995). Although these observational studies strongly suggest that close companion galaxies in QSO hosts play a vital role in triggering QSO activities (Bahcall et al. 1995), it is still theoretically unclear why QSO host galaxies so frequently have companions and how QSO activities are physically associated with the formation and the evolution of such companion galaxies. In this Letter, we numerically investigate both gas fueling to the seed black holes located in the central part of two disks in a gas-rich merger and the morphological evolution of the merger in order to present a plausible interpretation on the origin of small companion galaxies frequently observed in QSO host galaxies. We here demonstrate that the observed QSO companion galaxies are formed in the outer part of strong tidal tails during gas-rich major galaxy merging and then become self-gravitating compact galaxies orbiting elliptical galaxies formed by merging. We furthermore demonstrate that such companion galaxies are located within a few tens kpc of elliptical galaxies when efficient gas fueling to the central seed QSO black holes continues. We thus suggest that both the formation of QSO companion galaxies and the activation of QSO nucleus result from one physical process of gas-rich major galaxy merging. We furthermore discuss whether such companion galaxies formed in QSO hosts can finally become compact elliptical galaxies that are frequently observed in the present-day bright massive galaxies. \section{Model} We construct models of galaxy mergers between gas-rich disk galaxies with equal mass by using Fall-Efstathiou model (1980). The total mass and the size of a progenitor disk are $M_{\rm d}$ and $R_{\rm d}$, respectively. From now on, all the mass and length are measured in units of $M_{\rm d}$ and $R_{\rm d}$, respectively, unless specified. Velocity and time are measured in units of $v$ = $ (GM_{\rm d}/R_{\rm d})^{1/2}$ and $t_{\rm dyn}$ = $(R_{\rm d}^{3}/GM_{\rm d})^{1/2}$, respectively, where $G$ is the gravitational constant and assumed to be 1.0 in the present study. If we adopt $M_{\rm d}$ = 6.0 $\times$ $10^{10}$ $ \rm M_{\odot}$ and $R_{\rm d}$ = 17.5 kpc as a fiducial value, then $v$ = 1.21 $\times$ $10^{2}$ km/s and $t_{\rm dyn}$ = 1.41 $\times$ $10^{8}$ yr, respectively. In the present model, the rotation curve becomes nearly flat at 0.35 $R_{\rm d}$ with the maximum rotational velocity $v_{\rm m}$ = 1.8 in our units. The corresponding total mass $M_{\rm t}$ and halo mass $M_{\rm h}$ are 5.0 and 4.0 in our units, respectively. The radial ($R$) and vertical ($Z$) density profile of a disk are assumed to be proportional to $\exp (-R/R_{0}) $ with scale length $R_{0}$ = 0.2 and to ${\rm sech}^2 (Z/Z_{0})$ with scale length $Z_{0}$ = 0.04 in our units, respectively. The Toomre's parameter (\cite{bt87}) for the initial disks is set to be 1.2. The collisional and dissipative nature of the interstellar medium is modeled by the sticky particle method (\cite{sch81}). Star formation is modeled by converting the collisional gas particles into collisionless new stellar particles according to the Schmidt law (Schmidt 1959) with the exponent of 2.0. The initial gas mass fraction ($f_{\rm g}$) is considered to be a free parameter ranging from 0.1 (corresponding to a gas poor disk) to 0.5 (a very gas-rich one). We here present the result of the model with $f_{\rm g}=0.5$, because this model most clearly shows the typical behavior of QSO companion formation. The dependence of the details of QSO companion formation on $f_{\rm g}$ will be described by our future paper (Bekki 1999). The orbital plane of a galaxy merger is assumed to be the same as $xy$ plane and the initial distance between the center of mass of merger progenitor disks is 8.0 in our units (140 kpc). Two disks in the merger are assumed to encounter each other parabolically with the pericentric distance of 1.0 in our units (17.5 kpc). The intrinsic spin vector of one galaxy in a merger is exactly parallel with $z$ axis whereas that of the other is tilted by ${30}^{\circ}$ from $z$ axis. The present study describes the QSO companion formation only for a nearly prograde-retrograde merger in which only one intrinsic spin vector of a merger progenitor galaxy is nearly parallel with orbital spin vector of the merger. The dependence of the details of QSO companion formation processes on the initial orbital configurations of galaxy mergers will be given by Bekki (1999). The number of particles used in a simulation is 20000 for dark halo components, 20000 for stellar ones, and 20000 for gaseous ones. All the calculations including the dissipative and dissipationless dynamics and star formation have been carried out on the GRAPE board (\cite{sug90}) at Astronomical Institute of Tohoku University. The parameter of gravitational softening is set to be fixed at 0.03 in all the simulations. By using this merger model, we firstly investigate morphological and dynamical evolution of a gas-rich major galaxy merger with a particular emphasis on the formation of close small companions (dwarf-like galaxy) in the merger. Secondly, we investigate when and how QSO activities are triggered by major galaxy merging by counting total mass of interstellar gas accumulated within the central 100 pc of a galaxy merger. In order to estimate the gas mass in a explicitly self-consistent manner, we initially place a collisionless particle with the mass equal to $3.0 \times {10}^{6}$ in the mass center of a disk and regard this particle as a `seed black hole'. We then investigate both the time evolution of the orbit of the seed black hole and the total gas mass transferred to the central 100 pc around the black hole. Here we hypothetically assume that interstellar gas transferred to the central 100 pc around the seed black hole can be furthermore fueled to the central sub-pc region where a massive black hole gravitationally dominates and utilizes gas falling onto the accretion disk for a QSO activity. The reason for our adopting this assumption is that we regard a certain mechanism for gas fueling to the sub-pc region, such as the so-called `bars within bars' proposed by Shlosman, Frank, \& Begelman (1989), as being occurred naturally in the high-density self-gravitating central regions of galaxy mergers. The above two-fold investigation just allows us to address questions as to when and how galaxy merging not only forms small companions but also triggers QSO nuclear activities. \placefigure{fig-1} \placefigure{fig-2} \section{Result} Figure 1 describes how a QSO companion galaxy is formed by gas-rich major galaxy merging. As two gas-rich disks merge to form a tidal tail composed of gas and stars (the time $T$ = 1.1 Gyr), the stellar components in the tail first collapse to form a self-gravitating dwarf-like object. Gaseous components are then swept into the deep gravitational potential well of the dwarf galaxy to form a massive gaseous clump owing to the enhanced gaseous dissipation in the shocked region of the tidal tail and the dwarf. Star formation proceeds very efficiently in the high density gas clump, and consequently new stellar components are formed in the dwarf galaxy ($T$ = 1.7 Gyr). The physical processes of the dwarf galaxy formation in the present star-forming galaxy merger are essentially the same as those described by Barnes \& Hernquist (1992). This self-gravitating dwarf galaxy can then orbit an elliptical galaxy formed by galaxy merging without significant radial orbital decay due to dynamical friction between the dwarf and the host elliptical and tidal destruction by the elliptical ($T$ = 2.3 and 2.8 Gyr). Total mass in the dwarf at $T$ = 2.3 Gyr is roughly estimated to be $\sim 2.7 \times {10}^{9} {\rm M}_{\odot}$ corresponding to 4.5 \% of the initial disk mass. The gas mass fraction of the dwarf is rather large ($\sim 25\%$), which reflects the fact of the dwarf's being formed in the gas-rich tidal tail. About 45 \% of stellar components of the dwarf are very young stars formed from gaseous components of the tidal tail, which implies that this dwarf galaxy can be observed to show very blue colors until its hot and massive stars died out. Considering that the present gas-rich star-forming merger model also shows efficient gas fueling to the central seed black holes (as is described later), we regard the above results as demonstrating clearly that the dwarf galaxy formed in galaxy merging can be observed as a companion galaxy in a QSO host galaxy. Figure 2 shows the star formation history of the merger and the time evolution of gas mass located within the central 100pc around the seed black holes of the merger. Star formation rate becomes maximum ($\sim 378 \rm M_{\odot}/\rm yr$) at $T$ = 1.3 Gyr, when two disks finally merge to form an elliptical galaxy and the efficient redistribution of angular momentum and gaseous dissipation by cloud-cloud collisions cooperate to form the extremely high-density gaseous regions in the central part of the merger. After the intense secondary starburst, the star formation then rapidly declines owing to the efficient gas consumption by the starburst. Gas fueling to the central seed black holes becomes maximum ($6.5 \times {10}^{8} \rm M_{\odot}$) at $T$ = 1.3 Gyr, which is the same as the maximum starburst of the merger. Gas supply for the seed black holes is greatly controlled by the rapid gas consumption by star formation, and consequently gas fueling gradually declines after the completion of the secondary starburst. The gas fueling in the present study tends to be more efficient in the late phase of galaxy merging ($ T > 1.3 $ Gyr) than in the early one ($ T < 1.3 $ Gyr). Assuming that all of the gas transferred to the central 100pc around the seed black holes can be directly accreted onto the accretion disk of the black holes, we can estimate that the mean accretion rate in the merger late phase (1.3 Gyr $<T<$ 2.3 Gyr) is $6.3 \rm M_{\odot}/\rm yr$. The derived accretion rate is sufficient enough to trigger the typical magnitude of QSO activity (e.g., Rees 1984). These results imply that secondary massive starburst and QSO nuclear activity (AGN) can be observed to coexist in a QSO host galaxy, which is consistent with the observational evidence that some of QSO host galaxies show very bluer colors and spectroscopic properties indicative of the past starburst (MacKenty \& Stockton 1984; Boronson, Perrson, \& Oke 1985; Stockton \& Rigeway 1991). Thus Figure 1 and 2 clearly demonstrate that gas-rich major galaxy merging not only contributes to the formation of a companion galaxy orbiting a merger remnant but also triggers QSO nuclear activities. Accordingly our numerical study can naturally explain why QSO host galaxies, some of which are actually observed to be ongoing mergers and elliptical galaxies (e.g., Bahcall et al. 1997), are more likely to have close small companion galaxies; This is essentially because both QSO host galaxies with pronounced nuclear activities and their companions result from $one$ physical process of major galaxy merging. Our numerical studies furthermore provide the following three predictions on physical properties of QSO companions and hosts. First prediction is that the luminosity of a QSO companion galaxy is roughly proportional to that of the QSO host, principally because the mass of tidal debris that is a progenitor of a QSO companion depends strongly on the initial mass of a galaxy merger. Second is that a QSO companion has very young stellar population formed in secondary starburst of galaxy merging and thus shows photometric and spectroscopic properties indicative of starburst or post-starburst. Third is that not all of galaxy mergers can create QSO companions galaxies, essentially because nearly retrograde-retrograde mergers can not produce strong tidal tails indispensable for the formation of companion galaxies because of the weaker tidal perturbation of the mergers (The details of the physical conditions required for the formation of QSO companions will be described in Bekki (1999)). We suggest that future observational studies on the dependence of the luminosity-ratio of QSO hosts to QSO companions on QSO host luminosity, age and metallicity distribution of stellar populations of QSO companions, and the probability that QSO host galaxies have companion galaxies physically associated with them can verify the above three predictions and thereby can determine whether major galaxy merging is a really plausible model of QSO companion formation. \section{Discussion and Conclusion} The fate of QSO companion galaxies is an interesting problem of the present merger scenario of QSO companion formation. We here propose that some of the companions finally evolve into compact elliptical galaxies (cE) that have typical blue magnitude $M_{\rm B}$ ranging from -18 mag to -14 mag, truncated de Vaucouleurs luminosity profile, color-magnitude relation of giant ellipticals, typically solar-metallicity, and higher degree of global rotation (Faber 1973; Wirth \& Gallagher 1984; Nieto \& Prugniel 1987; Freedman 1989; Bender \& Nieto 1990; Burkert 1994). The essential reason of this proposal is described as follows. Burkert (1994) numerically investigated the dynamical evolution of proto-galaxies experiencing an initial strong starburst and the subsequent violent relaxation in the tidal external gravitational field of a massive elliptical galaxy and revealed that the observed peculiar properties of cEs are due to the external tidal field around progenitor proto-galaxies of cEs. Burkert (1994) accordingly proposed a scenario in which satellite proto-galaxies revolving initially around a bright elliptical galaxy eventually form cEs after violent cold collapse and strong starburst around the galaxy. Although his model of cE formation is not directly related to physical processes of gas-rich major galaxy merging, the physical environment of cE formation in his model is very similar to that of gas-rich galaxy merging; Tidal debris collapses to form a self-gravitating small galaxy in the rapidly changing external gravitational field of two merging disk galaxies in the present study. Accordingly it is not unreasonable to consider that some of companion galaxies created in tidal tails finally become cEs orbiting elliptical galaxies formed by major galaxy merging. The observational fact that cEs exist almost exclusively as satellites of bright massive galaxies (Faber 1973; Burkert 1994) strengthens the validity of the proposed evolutionary link between QSO companions and cEs. Furthermore, the larger degree of global rotation in kinematics observed in cEs (e.g., Bender \& Nieto 1990) seems to be consistent with the proposed scenario, since QSO companions are created in the tidal debris of rotationally supported disk galaxies in the scenario. The present numerical study unfortunately cannot investigate in detail structural and kinematical properties of companion galaxies formed in galaxy mergers because of very small particle number of the simulated companion ($\sim 800$). Our future high resolution simulations with the total particle number of $\sim {10}^{7}$ will enable us to compare the numerical results of structural, kinematical, and chemical properties of QSO companions formed in major mergers with observational ones of cEs located near giant ellipticals in an explicitly self-consistent manner and thereby answer the question as to the evolutionary link between intermediate and high redshift QSO companions and the present-day cEs. The most important observational test to assess the validity of the proposed formation scenario of QSO companion galaxies is to investigate whether a QSO companion galaxy has younger stellar populations formed by secondary starburst and thus shows photometric and spectroscopic properties indicative of starburst or post-starburst. Canalizo \& Stockton (1997) recently investigated spectroscopic properties of companion galaxies in three QSOs (3CR 323.1, PG 1700+518, PKS 2135-147) and found that the spectra of a companion galaxy in QSO PG 1700+518 shows both strong Balmer absorption lines from a relatively young stellar population and Mg I $b$ absorption feature and the 4000 $ \rm \AA $ break from an old stelar population. Stockton, Canalizo, \& Close (1998) furthermore demonstrated that the time that has elapsed since the end of the most recent major starburst event in the companion of QSO PG 1700+518 is roughly 0.085 Gyr, based on the spectral energy distribution derived from adaptive-optics image in $J$ and $H$ band. These observational results on the post-starburst signature of QSO companions are consistent reasonably well with the proposed scenario which predicts that a QSO companion galaxy contains both relatively old stellar populations previously located in merger progenitor disks and very younger stellar populations formed in gas-rich tidal tails. Detailed spectroscopic studies of QSO companion galaxies, such as Canalizo \& Stockton (1997) and Stockton, Canalizo, \& Close (1998), have not been yet so accumulated. Future extensive spectroscopic studies of companions in each of intermediate and high redshift QSOs will clarify the age distribution of stellar populations of the companions and thus determine whether most of QSO companions are really formed in major galaxy mergers. We conclude that gas-rich major galaxy merging can naturally explain the prevalence of small companion galaxies in QSO hosts; The essential reason for the origin of QSO companions is that strong tidal gravitational field of major galaxy merging both triggers the formation of companions and provides efficient fuel for QSO nuclear activities. This explanation of QSO companion formation is consistent reasonably well with the observational fact that QSO nucleus are already activated though the companions are still located in the vicinity of the QSO hosts ($\sim$ a few tens kpc from the center of the hosts). Our numerical simulations accordingly suggest that the observed companion galaxies in QSO hosts are not the direct $cause$ of QSO nuclear activities but the $result$ of gas-rich major galaxy merging. Although minor galaxy merging between small companion galaxies and giant elliptical galaxies or disk ones is demonstrated to be closely associated with secondary massive starburst in disks (Mihos \& Hernquist 1995) and strong starburst in shell galaxies (Hernquist \& Weil 1992), the present study implies that this minor merging is probably less important in the activation of QSO nucleus and the formation of QSO companions. The present study provides only one scenario of QSO companion formation, thus we lastly stress that physical processes related to the companion formation are likely to be more variously different and complicated than is described in the present study. \acknowledgments K.B. thank to the Japan Society for Promotion of Science (JSPS) Research Fellowships for Young Scientist. \newpage
\section{Introduction} The discovery of kilohertz quasiperiodic oscillations (QPO's) in the low mass X-ray neutron star (NS) binaries (Strohmayer {\it et al.\/} 1996; Van der Klis {\it et al.\/} 1996 and Zhang {\it et al.\/} 1996) has stimulated both theoretical and observational studies of these sources. In the upper part of the spectrum (400- 1200 Hz) for most of these sources, two frequencies $\nu_k$ and $\nu_h$ have been seen. Initially, the fact that for some sources, the peak separation frequency $\Delta \nu=\nu_h-\nu_k$ does not change much led to the beat frequency interpretation (Strohmayer {\it et al.\/} 1996; Van der Klis 1998) which was presented as a concept for the first time in the paper by Alpar \& Shaham (1985). Beat-frequency models, where the peak separation is identified with the NS spin rate have been challenged by observations: for Sco X-1, $\Delta\nu$ varies by 40\% (van der Klis {\it et al.\/} 1997 hereafter VK97) and for source 4U 1608-52, $\Delta\nu$ varies by 26\% (Mendez {\it et al.\/} 1998). Mounting observational evidence that $\Delta\nu$ is not constant demands a new theoretical approach. For Sco X-1, in the lower part of the spectrum, VK97 identified two branches (presumably the first and second harmonics) with frequencies 45 and 90 Hz which slowly increase in frequency when $\nu_k$ and $\nu_h$ increase. Furthermore, in the spectra observed by Rossi X-ray Timing Explorer (RXTE) for 4U 1728-34, Ford and van der Klis (1998, herein FV98) found low frequency Lorentzian (LFL) oscillations with frequencies between 10 and 50 Hz. These frequencies as well as break frequency, $\nu_{break}$ of the power spectrum density (PSD) for the same source were shown to be correlated with $\nu_k$ and $\nu_h$. It is clear that the low and high parts of the PSD of the kHz QPO sources should be related within the framework of the same theory. Difficulties which the beat frequency model faces are amplified by the requirement of relating the observed low frequency features, described above, with $\nu_k$ and $\nu_h$. Recently, a different approach to this problem has been suggested: kHz QPO's in the NS binaries have been modeled by Osherovich \& Titarchuk (1999) as Keplerian oscillations in a rotating frame of reference. In this new model the fundamental frequency is the Keplerian frequency $\nu_k$ (the lower frequency of two kHz QPO's) \begin{equation} \nu_k={{1}\over{2\pi}}\left({{GM}\over{R^3}}\right)^{1/2}, \end{equation} where G is the gravitational constant, M is the NS mass, and R is the radius of the corresponding Keplerian orbit. The high QPO frequency $\nu_h$ is interpreted as the upper hybrid frequency of the Keplerian oscillator under the influence of the Coriolis force \begin{equation} \nu_h=[\nu_k^2+(\Omega/\pi)^2]^{1/2}, \end{equation} where $\Omega$ is the angular rotational frequency of the NS magnetosphere. For three sources (Sco X-1, 4U 1608-52 and 4U 1702-429), we demonstrated that the solid body rotation ($\Omega=\Omega_0=const$) is a good first order approximation. Slow variation of $\Omega$ as a function of $\nu_k$ within the second order approximation is related to the differential rotation of the magnetosphere controlled by a frozen-in magnetic structure. This model allows us to address the relation between the high and low frequency features in the PSD of the neutron systems. We interpreted the $\sim 45$ and $90$ Hz oscillations as 1st and 2nd harmonics of the lower branch of the Keplerian oscillations in the rotating frame of reference: \begin{equation} \nu_L=(\Omega/\pi)(\nu_k/\nu_h)\sin\delta, \end{equation} where $\delta$ is the angle between ${\bf \Omega}$ and the vector normal to the plane of the Keplerian oscillations. For Sco X-1, we found that the angle $\delta=5.5^o$ fits the observations. In this Letter we include the LFL oscillations and related break frequency phenomenon in our classification. We attribute LFL oscillations to radial oscillations in the viscous boundary layer surrounding a neutron star. According to the model of Shakura \& Sunyaev (1973, hereafter SS73), the innermost part of the Keplerian disk adjusts itself to the rotating central object (i.e. neutron star). The recent modelling by Titarchuk, Lapidus \& Muslimov (1998, hereafter TLM) led to the determination of the characteristic thickness of the viscous boundary layer $L$. In the following section, we present the extension of this work to relate the frequency of the viscous oscillations $\nu_v$ and $\nu_{break}$ with $\nu_k$. Comparison with the observations is carried out for 4U 1728-34. The last section of this Letter contains our theoretical classification of kHz QPO's and related low frequency phenomena. \section{Radial Oscillations and Diffusion in the Viscous Boundary Layer} We define the boundary layer as a transition region confined between the NS surface and the first Keplerian orbit. The radial motion in the disk is controlled by the friction and the angular momentum exchange between adjacent layers resulting in the loss of the initial angular momentum by an accreting matter. The corresponding radial transport of the angular momentum in a disk is described by the equation (e.g. SS73): \begin{equation} \dot M {d\over {dR}}(\omega R^2) = 2\pi {d\over {dR}} (W_{r\varphi}R^2), \end{equation} where $\dot{M}$ is the accretion rate, and $ W_{r\varphi}$ is the component of a viscous stress tensor which is related to the gradient of the rotational frequency $\omega$, namely \begin{equation} W_{r\varphi}=-2\eta HR{{d\omega}\over{dR}}, \end{equation} where $H$ is a half-thickness of a disk, and $\eta$ is the turbulent viscosity. The nondimensional parameter which is essential for equation (4) is the Reynolds number for the accretion flow \begin{equation} \gamma={{\dot M}\over{4\pi\eta H}}={{3R v_r}\over {{\it v}_t{\it l}_t}}, \end{equation} which is the inverse $\alpha-$parameter in the SS73-model; $v_r$ is a characteristic velocity, $v_t$ and $l_t$ are a turbulent velocity and related turbulent scale respectively. Equations $\rm \omega=\omega_0~{\rm~at}~R=R_0$ (NS radius) and ${\rm \omega=\omega_K~at~R=R_{out}}$ (radius where the boundary layer adjusts to the Keplerian motion), and $\rm {{d\omega}\over{dr}}={{d\omega_k}\over{dr}}~ at~\rm R=R_{out}$ were assumed by TLM as boundary conditions. Thus the profile $\omega(R)$ and the outer radius of the viscous boundary layer $R_{out}$ are uniquely determined by these boundary conditions. Presenting $\omega(R)$ in terms of dimensionless variables: namely angular velocity $\theta=\omega/\omega_0$, radius $ r=R/R_0$ ($ R_0=x_0R_s$, $ R_s=2GM/c^2$ is the Schwarzschild radius), and mass $ m=M/M_{\odot}$, we express Keplerian angular velocity as \begin{equation} \theta_K={{6}/(a_K r^{3/2}}), \end{equation} where $a_K=m(x_0/3)^{3/2}(\nu_0/363~{\rm Hz})$ and the NS rotational frequency $\nu_0$ has a particular value for each star. The particular coefficient, 6, presented in formula (7) is obtained for the frequency of nearly coherent (burst) oscillations for 4U 1728-34, i.e. for $\nu_0=363$ Hz. The solution of equations (4-5 ) satisfying the above boundary conditions is \begin{equation} \theta(r)=D_1 r^{-\gamma} + (1-D_1) r^{-2}, \end{equation} where $D_1=(\theta_{out}-r_{out}^{-2})/(r_{out}^{-\gamma} -r_{out}^{-2})$ and $\theta_{out}=\theta_K(r_{out})$. Equation $\theta^{\prime}(r_{out})=\theta_K^{\prime}(r_{out})$ determines $r_{out}$: \begin{equation} {3\over2} \theta_{out}=D_1 \gamma r_{out}^{-\gamma}+2(1-D_1)r_{out}^{-2}. \end{equation} The solution of equations (4-5) subject to the inner sub-Keplerian boundary condition has a regime corresponding to the super-Keplerian rotation (TLM). For such a regime matter piles up in the vertical direction thus disturbing the hydrostatic equilibrium. The vertical component of the gravitational force prevents this matter from further accumulation in a vertical direction and drives relaxation oscillations. The radiation drag force, which is proportional to the vertical velocity, determines the characteristic decay time of the vertical oscillations (TLM). The characteristic time $t_r$, over which the matter moves inward through this region, bounded between the innermost disk and relaxation oscillations zone is \begin{equation} t_r\sim {L\over v_r}, \end{equation} where $L=R_{out}-R_0$ is the characteristic thickness of this region. Even though the specific mechanism providing the modulation of the observed X-ray flux over this timescale needs to be understood, this timescale apparently ``controls'' the supply of accreting matter into the innermost region of the accretion disk. Any local perturbation in the transition region would propagate diffusively outward over a timescale \begin{equation} t_{diff}\sim \left({L\over {l_{fp}}}\right)^2 {{l_{fp}}\over v_r}, \end{equation} where $l_{fp}$ is the mean free path of a particle. Note, that the $\gamma-$parameter is proportional to the accretion rate (see Eq. 6), and therefore $v_r\propto \gamma$. Using this relationship, we can exclude $v_r$ from the above equations and get the relations for the corresponding inverse timescales (frequencies). For the frequency of viscous oscillations \begin{equation} \nu_{v}\propto {\gamma\over {r_{out}-r_0}}, \end{equation} and for the break frequency, related to the diffusion \begin{equation} \nu_{break}\propto{\gamma\over {(r_{out}-r_0)^2}}. \end{equation} In the following section, we compare the predictions of this model with the observations and also establish the theoretical relation between $\nu_v$ and $\nu_{break}$. \section{Comparisons with Observations} The results of FV98 for the low frequency Lorentzian in the X-ray binary 4U 1728-34 are presented in Figure 1 and for the break frequency $\nu_{break}$ in Figure 2. In Figure 1, crosses represent the frequencies (with the appropriate error bars) observed during four days. Data collected on February 16 (open circles) are situated apart from the rest of observations and they are not included in the empirical power law fit which is suggested by FV98. In the work discussed above, the authors plotted the observed low frequencies versus high-frequency QPO which for all days, except February 16 was $\nu_k$ and apparently for February 16 it was $\nu_h$. Our theoretical curve for $\nu_v$ versus $\nu_k$ is based on equation (12). The $\chi^2$ dependence on this parameter is rather strong: the parabola $\chi^2=38024-73076\cdot a_k+35732\cdot a_k^2$ has a minimum at $a_k=1.03$, which determines the best fit. Using $\Omega/2\pi=340$ Hz in the upper hybrid relation (2), we calculate $\nu_k$ for the points observed on February 16 and show that they belong to the set of frequencies modeled by our theoretical curve for the viscous radial oscillations (closed circles). Identification of the observed $\nu_{break}$ with the inverse diffusion time (formulas 11 and 13) is illustrated by theoretical curves in Figure 2. It is worth noting that these two correlations with kHz frequencies are fit by two theoretical curves using {\it only one parameter $a_k$}. The $\chi^2-$ dependence on $a_k$ is obtained with inclusion of all data points for the break and low frequency correlations (75 data points). The theoretical dependences of $\nu_k$ and $r_{out}$ on $\gamma-$parameter are calculated numerically using equations (7) and (9) and employed here for calculations of the theoretical curves in Figures 2 and 3 using equations (12) and (13). We were unable to interpret data for $\nu_{break}$ collected in February 16 (open circles). Neither $\nu_v$ nor $\nu_{break}$ in our theory have a power law relation with $\nu_k$. However, the theoretical relation between $\nu_{break}$ and $\nu_v$, shown in Figure 3 by a solid curve, is close to the straight line (in log-log diagram), suggesting an approximate power law \begin{equation} \nu_{break}=0.041\nu_v^{1.61}. \end{equation} This relation is derived from the theoretical dependence for the best fit parameter $a_k=1.03$. Observations of FV98 (except February 16) are also presented in the Figure 3. \centerline {\bf 4. Discussion and Conclusions} We present a model for the radial oscillations and diffusion in the viscous boundary layer surrounding the neutron star. Our dimensional analysis has identified the corresponding frequencies $\nu_v$ and $\nu_{break}$ which are consistent with the low Lorentzian and break frequencies for 4U 1728-34. and predicted values for $\nu_{break}$ related to the diffusion in the boundary layer are consistent with the break frequency observed for the same source. Both oscillations (Keplerian and radial) and diffusion in the viscous boundary layer are controlled by the same parameter - Reynolds number $\gamma$ which in turn is related to the accretion rate. It is shown in TLM that $\nu_k$ is a monotonic function of $\gamma$. Therefore, the observed range of $\nu_k$, (350-900 Hz) corresponds to the range $1<\gamma<5$ (or $0.2<\alpha<1$). The results in this Letter extend the classification of kHz QPO's and the related low frequency phenomena suggested by Osherovich \& Titarchuk 1999. Figure 4 summarizes the new classification. Solid lines represent our theoretical curves and open circles observations for Sco X-1 (from VK97). As one can see, formulas (2) and (3), for the Keplerian oscillator under the influence of the Coriolis force, reproduce the observations well. Indeed, $\Delta\nu=\nu_h-\nu_k$ is not constant, as observed (see OT99 for details of comparisons of the data with the theory). Effectively, the main viscous frequency $\nu_v$ and the diffusive $\nu_{break}$ introduce the second oscillator with two new branches in the lower part of the spectra. The unifying characteristic of spectra for both oscillators is the strong dependence on $\nu_k$. This common dependence on $\nu_k$ can be viewed as a result of the interaction between Keplerian oscillator and the viscous oscillator which share the common boundary at the outer edge of the viscous transition layer. Our parametric study indicates that the power law index 1.6 in Eq. (14) should be the same for different neutron stars. We expect a similar relation for black holes but with a distinctly different index. The found value of $a_k$ leads ultimately to independent constraints in the determination of mass and radius for the neutron star (Haberl \& Titarchuk 1995). LT thanks NASA for support under grants NAS-5-32484 and RXTE Guest Observing Program. The authors acknowledge discussions with Alex Muslimov, Jean Swank, Lorella Angelini, Will Zhang, Joe Fainberg and fruitful suggestions by the referee. Particularly, we are grateful to Eric Ford, and Michiel van der Klis, for the data which enable us to make comparisons with the data in detail.
\section{Introduction} \label{sec:Intro} This paper deals with phase space parameterizations of one-dimensional {\em billiard map} eigenfunctions for polygonal enclosures. Specifically, we shall deal with the Bargman-Husimi representation and study the distribution of its zeroes for regular, irregular and bouncing ball modes. Such a study has been carried out before for integrable and chaotic billiards \cite{tualle_voros,leb_vor_95} and these systems are now reasonably well understood in the sense that the distribution reflects a correspondence with the underlying classical dynamics. As with most other objects of interest in generic polygonal (pseudo-integrable) billiards, the distribution of zeroes is interesting if only to explore the existence of such a correspondence with the classical system. Of all possible Hamiltonian systems, billiards are perhaps the best understood category and exhibit the entire gamut of classical dynamics depending on the shape of the enclosure. Of these, polygonal billiards form an important sub-category and apart from the rectangle and the triangles $(\pi/3,\pi/3,\pi/3)$, $(\pi/2,\pi/3,\pi/6)$, $(\pi/2,\pi/4,\pi/4)$, all other polygonal enclosures are non-integrable \cite{pjr_mvb}. Further, the ones with rational interior angles are pseudo-integrable; they have two constants of motion as in integrable systems and yet their invariant surface in phase space has a genus, $g > 1$. One of the simplest examples of a pseudo-integrable system is the $\pi/3$ enclosure for which $g = 2$ {\it i.e. the invariant surface is a double torus}. Here, as in other pseudo-integrable billiards, an initial (parallel) beam of trajectories splits after successive encounters with the $2\pi/3$ (in general $m\pi/n, m > 1$) vertex and traverse different paths. There are several important consequences of pseudo-integrability at the classical level that are now known. However, as far as semiclassics is concerned, pseudo-integrable billiards are still rather poorly understood. When the dynamics is integrable, an EBK ansatz for the wavefunction \cite{keller} \begin{equation} \psi (q) \sim \sum_{j=1}^N A_j\exp(i S_j/\hbar) \label{eq:EBK} \end{equation} \noindent works well at least in the limit $\hbar \rightarrow 0$. In the above, $S_j$ are the (finitely many) branches of the classical action at energy $E$ and $A_j$ are constant amplitudes for integrable polygons. Such an ansatz however does not work for pseudo-integrable billiards even though the number of sheets that constitute the invariant surface is still finite. We shall not discuss the reasons for its breakdown here but merely remark that no definite behaviour for pseudo-integrable eigenfunctions is known. For classically chaotic systems on the other hand, the Schnirelman theorem \cite{schnirel} (suitable phase-space measures constructed from the eigenfunctions must tend towards the classical phase-space ergodic measure as $\hbar \to 0$) does provide a semiclassical constraint albeit in a measure theoretic sense. Besides, there exist results on the amplitude distribution and spatial correlation function which have been subject to tests \cite{mcdonald88}. Despite the absence of any such result for pseudo-integrable polygons, numerical studies \cite{db90} such as those for the amplitude distribution or nodal plots suggest that typical eigenfunctions are irregular and broadly speaking, there is little to distinguish them from the eigenfunctions in chaotic systems. In the present paper, we shall try to refine this existing body of knowledge and will employ for this purpose a phase-space representation of quantum mechanics, which is known to highlight certain semiclassical features for integrable and chaotic systems. Our results are empirical, based on extensive numerical studies and can be simply expressed as follows : the eigenfunctions of polygonal billiard as viewed in the Husimi representation tend to be irregular {\em but nevertheless contain a subtle signature of classical pseudo-integrability}. The paper is organized along the following lines. In section~\ref{sec:Formalism}, we briefly review the Husimi - Bargman representations and the results on random analytic functions. We introduce the systems that we shall study and the quantum map under consideration in section~\ref{sec:models}. This is followed by our numerical results on the Husmini function and the density of zeroes in section~\ref{sec:numerics}. Finally, correlations are discussed in section~\ref{sec:correlations} and our conclusions are summarized in section~\ref{sec:Conclusions}. \section{Phase Space Representations} \label{sec:Formalism} Phase space representations of quantum wavefunctions are best suited in semiclassical studies since the quantum dynamics (Heisenberg equation) then appears as an explicit deformation of the classical dynamics (Liouville equation) by shifting the analysis onto the density operator $\rho = | \psi > < \psi |$. In quantum mechanics however, the phase space representation of a state is not unique since operators ($\hat{q}, \hat{p} $ for instance ) may be ordered in various ways while having the same classical analog. A general expression for a quasi-probability distribution function may be expressed as \cite{ZFG} \begin{equation} \rho_{(\Omega )}(q,p,t) = {1\over (2\pi)^2}\int~d^2\xi~e^{i(\xi^*z^* + \xi z) \hbar }~{\rm Tr}~[ \Omega\{e^{-i\xi^*\hat{a}\dagger} e^{-i\xi\hat{a}} \} \hat{\rho} ] \label{eq:basic} \end{equation} \noindent where $\Omega$ refers to the ordering that is chosen. The Wigner distribution follows from a symmetric ordering of ($\hat{q}, \hat{p} $) which implies \begin{equation} \Omega\{e^{-i\xi^*\hat{a}\dagger} e^{-i\xi\hat{a}} \} = e^{-i\xi^*\hat{a}\dagger - i\xi\hat{a}} \label{eq:weyl} \end{equation} \noindent while the Husimi function is a result of anti-normal ordering \begin{equation} \Omega\{e^{-i\xi^*\hat{a}\dagger} e^{-i\xi\hat{a}} \} = e^{-i\xi\hat{a}}e^{-i\xi^*\hat{a}\dagger} \label{eq:antinormal} . \end{equation} \noindent Using Eqns.~(\ref{eq:basic}) and (\ref{eq:weyl}), the distribution function in the Wigner representation, $\rho_w(q,p;\hbar)$, for a pure state can be explicitly written as : \begin{equation} \rho_w(q,p;\hbar) = {1\over (2\pi\hbar)^d} \int < q - \eta / 2 | \psi > < \psi | q + \eta / 2 > e^{ip.\eta/\hbar} d\eta \label{eq:wigner} \end{equation} \noindent where $d$ is the degree of freedom of a dynamical system. Thus, the expectation of a dynamical variable $\hat{A}$ is represented as \begin{equation} {\rm Tr}~[~\hat{A}~| \psi > < \psi |~] = \int A_w(q,p) \rho_w(q,p)~dqdp \label{eq:trace} \end{equation} \noindent where \begin{equation} A_w(q,p) = \int < q - \eta / 2 |~\hat{A}~| q + \eta / 2 > e^{ip.\eta/\hbar} d\eta \end{equation} The Wigner function however takes positive as well as {\em negative} values and oscillates violently with a wavelength $\hbar$ in phase space. A coarse grained distribution function is thus preferred and the Husimi function, \begin{equation} \rho_h(q,p;\hbar) = {1\over (\pi\hbar)^d} \int \rho_w(q',p';\hbar) e^{-\sum_{i=1}^{N}[{(q_i-q'_i)^2 \over 2(\Delta q_i)^2} + {(p_i-p'_i)^2] \over 2(\Delta p_i)^2}]} dp'dq' \label{eq:husimi} \end{equation} \noindent is one such example which can be expressed as a smoothened Wigner function. In this case, the smoothening is achieved through the Gaussian centred at a phase space point $(q,p)$. In Eq.~(\ref{eq:husimi}) above, \begin{equation} \Delta q_i = \sqrt{ {\hbar \over 2} } \sigma_i,~~~~~\Delta p_i = \sqrt{ {\hbar \over 2} } {1 \over \sigma_i} \label{eq:uncertain} \end{equation} \noindent are the uncertainties in $q$ and $p$ respectively. Note that $\rho_h$ is merely a minimum-uncertainty (m.u.) state decomposition of the wavefunction $\psi$ and can be expressed as \begin{equation} \rho_h(q,p;\hbar) = { | < z | \psi > |^2 \over 2\pi\hbar } \label{eq:husimi_coh} \end{equation} \noindent where \begin{equation} | z > = e^{ {-|z|^2\over 2}} \sum_{n=0}^\infty {z^n\over \sqrt{n!}} | n > \end{equation} \noindent $ \{ | n > \} $ are the harmonic oscillator number states, $a^\dagger = (\sigma^{-1/2} \hat{q} - \imath \sigma^{1/2} \hat{p})/ (\sqrt{2\hbar})$ and $ z = (\sigma^{-1/2} q - \imath \sigma^{1/2} p)/ (\sqrt{2\hbar})$ with $\sigma > 0$. Note that $ < z | z > = 1 $ while $ < z | z' > \neq 0 $. Written explicitly for $1~-$ degree of freedom, \begin{equation} <x|z> = \left ( {1\over 2\pi (\Delta q)^2} \right )^{1/4} e^{ipx~-~ {(x - q)^2 \over 4 (\Delta q)^2}} \end{equation} \noindent which is the minimum uncertainty wavepacket whose Wigner transform is the Gaussian used in Eq.~(\ref{eq:husimi}). From Eq.~(\ref{eq:husimi_coh}), it is evident that $\rho_h$ takes only positive values. The minimum wavepackets, $ | z > $ and $ < z | $ are eigenfunctions of $\hat{a}$ and $\hat{a}^\dagger$ respectively with eigenvalues $z$ and $z^*$. Eq.~(\ref{eq:husimi_coh}) follows directly from eqns.~(\ref{eq:basic}) and (\ref{eq:antinormal}) using the expansion of the identity operator \begin{equation} \hat{{\rm I}} = \int d\mu(z)~|z><z| \end{equation} \noindent where $d\mu(z) = dqdp/(2\pi\hbar)$. If the system under consideration is ergodic, the Husimi density $ \{\rho_h^{n}\} $, corresponding to a sequence of eigenstates $ \{\psi_n(q)\} $ with eigenvalues $ E_n \rightarrow E$, almost always converges to the classical Liouville measure $\mu_E$ over the energy surface $\Sigma_E$. Thus, if $f(q,p)$ is any smooth observable, \begin{equation} \int~f(q,p)~\rho_h^{n}~dqdp \rightarrow \int_{\Sigma_E}~f(q,p)~ d\mu_E~~~{\rm as}~~~E \rightarrow E_n. \label{eq:ergodic} \end{equation} \noindent Schnirelman's theorem however allows an occasional exception (e.g. scarred state) and for this reason, a more appropriate description of non-integrable eigenfunctions is desirable. In 1990, Leboeuf and Voros \cite{le_vo_jphys_a_90} proposed that the zeroes of the Husimi function provide a minimal description of quantum states \cite{more_recent}. The first step in this direction is the coherent state ($\sigma = 1$) or Bargman representation, $ < z | \psi > $ of a state $ | \psi > $ which maps unitarily the standard Hilbert space onto the space of {\em entire} functions with finite Bargman norm \begin{equation} \parallel \psi \parallel = {1 \over 2\pi\hbar} \int_{R^2} | \psi(z) |^2 e^{-|z|^2 } dq dp. \label{eq:barg_norm} \end{equation} \noindent One can thus consider $\psi(z)$ as a phase phase representation of the wavevector $ | \psi >$. Note that the zeroes of the Bargman and Husimi functions are identical. The Bargman function however contains information about the phase (of the wavefunction) as well and is hence a more fundamental object. For the standard case when the phase space is a plane (the Weyl - Heisenberg group, $W_1$), \begin{equation} \psi(z) = e^{ {-|z|^2\over 2}} \sum_{n=0}^\infty {a_n\over \sqrt{n!}}~z^n ~~~~~~~~~~~~~~~~~~~~~(W_1) \label{eq:W_1} \end{equation} \noindent where $a_n$ are the expansion coefficients of $ | \psi > $ in terms of the harmonic oscillator number states. Similar results can be written down for the sphere (~$SU(2)$~) and the pseudo-sphere (~$SU(1,1)$~) \cite{perelomov,leboeuf_recent} though unlike the case of $W_1$ or $SU(1,1)$, the Bargman representation of $| \psi >$ for $SU(2)$ is finite reflecting the compactness of phase space. For Hamiltonian systems however, energy conservation does ensure that the manifold is compact so that Eq.~(\ref{eq:W_1}) has, in practice, only a finite number of terms. Clearly then, the Husimi-Bargman zeroes specify a state completely. It is evident that the distribution of the Husimi-Bargman zeroes depends on the distribution of the expansion coefficients $\vec{a} = (a_1,a_2,\ldots,a_n)$. For chaotic systems, it is natural to expect that the choice of an arbitrary basis (harmonic oscillator in this case) makes $\vec{a}$ point in any direction of Hilbert space with equal probability \cite{leboeuf_recent}. The only constraint then comes from normalization so that $\sum a_n^2 = 1$. For purposes of computing the distribution of zeroes, this is equivalent to the assumption that the coefficients are drawn from a Gaussian distribution \cite{kac} \begin{equation} D(\vec{a}) = {1\over (2\pi)^N}~e^ {- \sum_i {|a_i|^2 \over 2} } \label{eq:gauss} \end{equation} \noindent Eq.~(\ref{eq:W_1}) with the above distribution is referred to as a {\em random analytic function}. Random analytic functions (RAF) for various groups have been studied in some detail when the coefficients are complex \cite{leboeuf_recent,bogo_bohi_lebo, leb_shukla,hannay} corresponding to systems without time reversal symmetry. The results point to a universal behaviour. Thus, the density of zeroes is uniform with spacings of the order of $1/\sqrt N$ and the 2-point correlation has a simple form \cite{hannay,hannay_all} independent of the location of the zeores. Importantly, random analytic functions do seem to model chaotic systems very well \cite{leb_shukla,shukla}. For RAF with real coefficients (systems with time reversal symmetry), Prosen \cite{prosen} has studied the density and the k-point correlations. The density in this case is non-uniform due to the presence of zeroes on the symmetry axis (the real line). Away from the real axis however, the density becomes uniform and in this region, correlations tend towards the case with complex coefficients. There are few numerical studies however on chaotic systems with time reversal symmetry though it might be expected that RAF with real coefficients do model them rather well. In contrast, it is known \cite{voros_pra} that for integrable systems, eigenfunctions follow a WKB-type Ansatz (see eq.~\ref{eq:EBK} ) in the Bargmann representation too, from which it follows that the zeroes lie on fixed curves which are anti-Stokes lines of the complex classical action in the $z$ variable, along which the zeros are equi-spaced with the separation of order $1/N$. For the sake of completeness, it may also be noted that a random polynomial \begin{equation} \psi(z) = a_0 + a_1z + a_2Z^2 + \ldots + a_Nz^N \label{eq:ranpoly} \end{equation} \noindent with coefficients distributed according to Eq.~(\ref{eq:gauss}), has zeroes which tend to accumulate around the unit circle \cite{bogo_bohi_lebo}. With this background, we shall explore the distribution of Husimi zeroes for polygonal billiard eigenfunctions in the following sections. Unless otherwise stated, we shall consider enclosures with unit perimeter and $\sigma = 1$ (coherent state). We shall also consider the energy, $E = 1$ and instead quantize $\hbar$ so that $\hbar = 1/k$. The $\hbar \rightarrow 0$ then corresponds to the classical dynamics at $E = 1$. \section{Polygonal Billiards and the Quantum Map} \label{sec:models} Classical billiards are enclosures within which a point particle undergoes specular reflection. The dynamics thus depends on its shape. For rational polygonal enclosures, the dynamics is constrained by two constants of motion such that the invariant surface is two-dimensional. For the rectangle and the integrable triangles, this is a torus for which $g=1$. For all other rational polygons, the invariant surface is topologically equivalent to a sphere with multiple holes ($g > 1$). The simplest example is a double torus (g = 2) which corresponds to enclosures such as the $\pi/3$ rhombus or the L-shaped billiard. In general, the genus of any rational polygon can be calculated from its interior angles. Thus, if $m_i\pi/n_i$ are the interior angles of a rational polygon, \begin{equation} g = 1 + {N\over 2} \sum_i {m_i - 1\over n_i} \label{eq:genus} \end{equation} \noindent where $N$ is the least common multiple of $n_i$ so that the number of sheets that constitute the invariant surface is $2N$. Thus various sets of internal angles may have the same genus but with different $N$ such that the number of distinct momenta spanned by a generic trajectory varies from enclosure to enclosure. While the genus does affect certain classical features of the system \cite{db_pramana}, its influence on quantum states is not known for certain. Studies on irrational and rational rhombus billiards show that there is little difference between the morphologies of generic eigenfunctions or their Husimi densities \cite{shudo_shimizu_95}. Shudo and Shimizu \cite{shudo_shimizu_95} even note that ``~\ldots the difference between random features of eigenfunctions of quantum polygonal and the desymmetrized dispersing system are minute \ldots ''. The only difference, they noted, was the occurrence of bouncing ball states though these can be observed in other chaotic systems such as the Stadium billiard. Our investigation of polygonal billiard eigenfunctions lies in this backdrop. Instead of the Husimi densities themselves, we shall study their zeroes following Tualle and Voros \cite{tualle_voros}. The systems we choose are triangles and rhombus billiards and for all practical purposes, these can be treated as pseudo-integrable systems irrespective of the internal angle \cite{see_hobson,high_genus}. The eigenvalues and eigenfunctions can be obtained by solving the Helmholtz equation \begin{equation} (\nabla^2 + E) \Psi(q) = 0 \label{eq:helm} \end{equation} \noindent with $\Psi(q) = 0$ on the boundary. The problem can however be reduced to an eigenvalue problem for an integral operator $K$ or a {\em Quantum Poincare Map} in various ways \cite{boasman} : \begin{eqnarray} \psi(s) & = & \oint ds' \psi(s') K_D(s,s';k) \\ K_D(s,s';k) & = & - {\imath k \over 2} \cos \theta(s,s') H_1^{(1)} (k|\vec{s} - \vec{s'}|) \\ \cos \theta(s,s') & = & \hat{n}(\vec{s}).\hat{\rho}(s,s') \label{eq:bim} \end{eqnarray} \noindent where $E = k^2$, $\hat{\rho}(s,s') = (\vec{s} - \vec{s'})/|\vec{s} - \vec{s'}|$ and $ \hat{n}(\vec{s})$ is the outward normal at the point $\vec{s}$. The unknown function is now the normal derivative on the boundary \begin{equation} \psi(s) = \hat{n}(\vec{s}).\nabla \Psi(\vec{s}) \label{eq:psi_Psi} \end{equation} \noindent and the full interior eigenfunction can be recovered through the mapping \begin{equation} \Psi(q) = - {\imath \over 4} \oint ds H_0^{(1)}(k|\vec{s} - \vec{s'}|) \psi(s) \label{eq:Psi_psi} \end{equation} \noindent Thus, the essential dynamical information lies within the reduced $1-d$ function $\psi(q)$ and we shall use this to study phase space representations and look at their zeroes. For an enclosure of unit perimeter (which we shall assume from now on) $ \psi(q + 1) = \psi(q)$. The Bargman transform, $\psi(z)$ thus obeys a quasi-periodicity condition as well \cite{tualle_voros} : \begin{equation} \psi(z + 1) = e^{{i \over \hbar}p}~\psi(z) \end{equation} \noindent and the norm-finiteness condition becomes : \begin{equation} \parallel \psi \parallel = {1 \over 2\pi\hbar} \int_{-\infty}^{+\infty} ~dp \int_0^1~dq~|\psi(z)|^2 e^{-|z|^2}~~<~~\infty \end{equation} \section{Husimi Zeroes in Polygons - Results} \label{sec:numerics} The distribution of Husimi-Bargman zeroes in polygonal billiards has not been investigated before and as remarked earlier, the only properties known about the eigenfunctions are from numerical studies. The lack of concrete results leaves us with little expectation and perhaps the only conjecture that can be made is that the distribution of Husimi zeroes of polygonal billiards should differ from the regularly spaced zeroes along fixed curves typical of integrable systems. Note that classical Poincare section plots in suitable (Birkhoff) co-ordinates do not immediately reveal the dramatic difference between integrable and pseudo-integrable polygons. In both cases, the points lie along a finite number of $\sin \theta = {\rm constant}$ lines where $\theta$ is the angle between the ray and the inward normal at the boundary point $q$. Thus there is little difference between the Poincare sections of the equilateral triangle and the $\pi/3$ rhombus. With increasing genus however, the number of such lines generally increase as the trajectory explores larger number of momentum directions. Semiclassically, the Husimi eigen-distribution function is known to be localized near the torus for integrable systems \cite{takahashi} while its zeroes distribute themselves along curves maximally distant from the invariant curves (anti-Stokes lines). As an example, we first consider the equilateral triangle billiard. Fig.~1 shows the Husimi distribution of a typical eigenstate with quantum number $(m,n) = (26,81)$ while Fig.~2 is a plot of its zeroes. Clearly, the Husimi distribution is peaked on the corresponding torus as evident from Fig.~3 while the zeroes lie on lines located away from the torus. Further, the zeroes are equispaced on each line though the spacings typically do vary from line to line. The zeroes do not always distribute themselves along straight lines in all integrable polygons and the equilateral triangle with its high symmetry is a rather special case. In fact, the distribution of zeroes of an equilateral state viewed in another enclosure (~related by symmetry - for instance the ($\pi/6,\pi/3,\pi/2$) triangle or the $\pi/3$ rhombus~) looks very different. Fig.~4 is an example where the fixed curves are not always straight lines though the zeroes are equi-spaced along each curve. As examples of pseudo-integrable polygons, we shall consider rhombus and triangle billiards. Since the choice of enclosure plays an important role in determining the distribution of zeroes, we shall use the $\pi/3$ rhombus to compare the regular and irregular states. Note that the regular states in this case correspond to equilateral triangle modes which vanish on the shorter diagonal and they comprise approximately half the total number of states in the $\pi/3$ enclosure (fig.~4 is an exmaple). The irregular states on the other hand are ``pure rhombus'' modes \cite{db90} which do not vanish on the shorter diagonal. Barring the bouncing-ball modes, ``pure rhombus'' modes display features typical of irregular wavefunctions. We shall look for the differences in the distribution of zeroes between (i) regular and irregular modes and (ii) bouncing-ball and non-bouncing-ball ``pure rhombus'' modes. Fig.~5 displays the zeroes of a typical irregular ``pure rhombus'' mode. The zeroes are no longer distributed along curves and they tend to diffuse all over the phase space. Note that there is a reflection symmetry in this case about the $q=0.25, 0.5$ and $0.75$ lines so the zeroes need only be viewed in a quarter of the phase space. Clearly, they are more or less randomly distributed with no clear alignment along any curve barring some exceptions where two or more zeroes are distributed around some $p = {\rm constant}$ line. These observations are in sharp contrast to the distribution of zeroes for integrable polygons. We next look at the zeroes of a neighbouring bouncing ball state. Studies on the stadium billiard have shown that the Husimi zeroes of bouncing ball modes are distributed randomly over the entire phase space as in case of irregular modes - an observation that may seem counter-intuitive keeping in mind the existence of approximate quantum numbers in the description of such states \cite{bai_hose_etal}. Fig.~6 shows the Husimi zeroes of a typical bouncing ball mode in the $\pi/3$ rhombus. The distribution is no different from the earlier case with few zeroes distributed around $p = {\rm constant}$ lines and the other zeroes distributed randomly. The symmetry of the rhombus leads to redundant zeroes and hence poor statistics as compared to an unsymmetric polygon at the same energy. However, it does show that the Husimi zeroes do not align themselves along fixed curves but rather tend to diffuse over phase space with some amount of clustering around a few $p =$ constant lines. As further evidence, we display the Husimi zeroes of a typical state in the ($\pi/4,\pi/5$) triangle in Fig.~7. They are indeed distributed over the entire classical phase space while the dashed lines indicate a tendency to cluster around certain momenta. This effect however seems to be pronounced only in systems with low genus. Thus for the triangle with internal angles ($97\pi/301,79\pi/501$), there seems to be little or no clustering (see Fig.~8) and the zeroes seem to be genuinely distributed over the entire phase space as in chaotic billiards. Fig.~9 shows a set of four histograms which illustrate this difference in clustering. The $x$ axis of the histograms give the momenta value and the $y$ axis shows the fraction of zeros occurring in a bin. The peaked distribution at specific $p$ values for the low genus ($\pi/4,\pi/5$) triangle indicates a clustering of its zeros. In contrast the high genus case shows an almost uniform distribution of zeros away from the real axis marked by a nearly flat histogram (barring the enhanced density around $p=0$). Thus, eigenstates of generic \cite{high_genus} pseudo-integrable billiards tend to behave like their chaotic counterparts insofar as the distribution of zeroes is concerned. This suggests that there is no obvious semiclassical correspondence in non-integrable polygonal billiards. In integrable polygons however, the correspondence is clear at least when $ \Delta q_i = \Delta p_i = \sqrt{\hbar/2} $ (see Eq.~(\ref{eq:uncertain}). However, when this not so (the minimum uncertainty state is not a coherent state \cite{coherent}), the zeroes tend to move with $\sigma$. As an example, we display here the zeroes of an equilateral triangle mode for two values of $\sigma$ in Fig.~10. When $\sigma = \sqrt{\hbar/2}$, the zeroes are equi-spaced and lie on a line. However, as $\sigma$ is reduced, the zeroes move outwards and realign themselves on a curve as shown in the figure. Finally, as $\sigma$ is reduced further, the zeroes start moving out of the classical phase space \section{Correlations} \label{sec:correlations} In the previous section, we found that the zeroes in non-integrable polygonal enclosures are uniformly distributed away from the real axis and hence are like those of random analytic functions with real coefficients which presumably model chaotic systems with time reversal symmetry. To ascertain how close the distributions are, we shall study here the nearest neighbour spacings distribution, $P(s)$ and the 2-point correlation, $R_2(r)$. \subsection{Nearest Neighbour Distribution} The nearest neighbour spacings distribution is the simplest statistic to perform though there exists no analytic predictions for RAF with real or complex coefficients. The curve in Fig.~11 for random analytic function is thus determined numerically. A total of approximately 25,000 zeroes from 50 eigenstates of three different non-integrable triangles has been used for computing the nearest neighbour distribution of generic polygons. The zeroes have been unfolded such that $\int~s~P(s)~ds = 1$ Fig.~11 shows a plot of the integrated spacings distribution, $I(s) = \int_0^s~P(s') ds'$ for polygons and a comparison with random analytic function having real coefficients. The agreement is fair but there are deviations indicating perhaps that the underlying assumption about the distribution of coefficients (see Eq.~(\ref{eq:gauss}) is not fully justified. Remarkably however, the Ginibre ensemble \cite{mehta,haake} of complex random matrices shows much better agreement as evident from fig.~11. In this case, the integrated spacings distribution \cite{haake}, $I_G(s) = i(<s> s)$ where $<s> = \int_0^\infty ds [ 1 - i(s) ] = 1.142929$ and \begin{equation} i(s) = 1 - \lim_{N \rightarrow \infty} \prod_{n=1}^{N-1}~[ e_n(s^2)~e^{-s^2} ] \end{equation} \noindent where \begin{equation} e_n(x) = 1 + {x \over 1!} + {x^2 \over 2!} + \ldots + {x^n \over n!} \end{equation} \noindent At small values of $s$, $I_G(s) \sim s^4$ and hence $P(s) \sim s^3$. In comparison, the nearest neighbour spacing distribution for uncorrelated points thrown at random on the plane exhibits no level repulsion. \subsection{Two-point correlation} For $SU(2)$ random analytic functions with complex coefficients, the k-point correlation function has been computed by Hannay analytically. In particular, the 2-point function, $R_2({\bf r_1},{\bf r_2}) = <~\rho({\bf r_1})~\rho({\bf r_2})~>$ depends only on the relative distance $r$ between points ${\bf r_1}$ and ${\bf r_2}$ since the density if uniform. In the asymptotic (number of zeroes, $N \rightarrow \infty$) limit \begin{equation} R_2(r) \simeq { (~\sinh^2~v + v^2~)~\cosh~v - 2v~\sinh~v \over \sinh^3~v } \label{eq:R_2_complex} \end{equation} \noindent where $v = \pi r^2/2$ and $r$ is measured in terms of the mean spacing ( $\sqrt{4\pi/N}$ for the sphere ). This result holds for other phase space topologies as well when $N \rightarrow \infty$ and the coefficients are complex. For systems with time reversal symmetry (real coefficients), the density is not uniform everywhere and hence $R_2({\bf r_1},{\bf r_2})$ is sensitive to the location of the zeroes. Away from the real axis however, $R_2$ has the limiting behaviour given by Eq.~(\ref{eq:R_2_complex}). For the Ginibre ensemble of complex random matrices, the density is uniform and the two point correlation (in unfolded units) \begin{equation} R_2({\bf r_1},{\bf r_2}) = 1 - \exp(- \pi\left| r_1 - r_2 \right|^2) \label{eq:R_2_ginibre} \end{equation} \noindent is a function of the distance between the two zeroes. Note that Eq.~(\ref{eq:R_2_ginibre}) does not have the characteristic hump at $r \simeq 1$ associated with random analytic functions. In Fig.~12, we present results for three different triangles. The close agreement suggests that there is possibly a universality in the distribution of zeroes of non-integrable polygons (corroborated by similar studies on the nearest neighbour). We next compare the average of the combined data with the predictions for the Ginibre ensemble (~see Eq.~(\ref{eq:R_2_ginibre})~) and Eq.~(\ref{eq:R_2_complex}). The deviations from the RAF predictions~\cite{fnote_r2} are evident while the Ginibre ensemble result agrees with our data very well. \section{Conclusions} \label{sec:Conclusions} We have studied the distribution of Husimi zeroes in polygonal billiards in this paper and our observations can be summarized as follows : \vskip 0.1 in $\bullet$ In integrable enclosures, the Husimi density is peaked on the classical torus and the zeroes lie equi-spaced on fixed curves that are located away from the torus when the minimum uncertainty state is a coherent state. \vskip 0.1 in $\bullet$ The zeroes tend to move as the uncertainties in position and momentum are varied even as they obey the minimum uncertainty relation. Thus, coherent states ($\Delta p = \Delta q = \sqrt{\hbar/2} $ ) are the most classical of all minimum uncertainty states. \vskip 0.1 in $\bullet$ A weak signature of pseudo-integrability can be associated with the clustering of some zeroes around a few lines as observed in some low-genus polygons. \vskip 0.1 in $\bullet$ For generic pseudointegrable enclosures, the zeroes tend to be randomly distributed over the entire phase space as in chaotic billiards or random analytic functions with real coefficients. This is especially true for polygons with high genus. \vskip 0.1 in $\bullet$ The nearest neighbour spacings distribution of zeroes and the two point correlation, $R_2(r)$, suggests that for pseudo-integrable billiards, the correlations are very well described by the Ginibre ensemble of complex random matrices. It is however not clear why this is so and a proper understanding is desirable. \section{Acknowledgements} The authors acknowledge stimulating discussions with Prof. A.~Voros and thank Dr. Pragya Shukla for valuable help in our studies on correlations. D.B also acknowledges several useful discussions on quasi probabilty distributions with Dr. R.~R.~Puri.
\section{Introduction} \vspace{-.15cm} An important part of the $B$ factory program at SLAC and KEK will be to search for CP violation in the mixing and decay of neutral $B$ mesons. In close analogy with the kaon system, the weak interaction allows mixing of the $B^0$ and the $\overline{B}^0$ through a second-order $\Delta B=2$ transition. The CP violation in this picture results from the interference between the amplitude for the decay $B^0 \rightarrow f$ to some CP eigenstate $f$ and the amplitude for mixing to occur first and then the decay, $B^0 \Rightarrow \overline{B}^0 \rightarrow f$. When the two amplitudes have a relative weak phase, CP violation results. Such weak phases arise in the Standard Model because of the complex CKM matrix \cite{CKM}. In the Wolfenstein parameterization of the CKM matrix\cite{Wolfenstein}, the phase of $V_{ub}$ is $\tan^{-1}(\eta/\rho)$ and both the CKM elements $V_{ub}$ and $V_{td}$ are expected to have large phases. Thus it is hoped CP violation may be observed in the $B$ system. Figure 1 shows schematically the current experimental bounds\cite{stone} on the CKM phase $(\rho,\eta)$ from $B_d^0$ mixing, from measurements of $V_{ub}/V_{cb}$ and from limits on $B_s$ mixing. Also shown is the bound from measurements of $\varepsilon_{\mbox{K}}$ in neutral kaons. Overlaid on the figure is a triangle whose sides are related to products of CKM elements that results from the requirement that the CKM matrix is unitary~\cite{jarlskog}: $V_{ub}V^*_{ud}+ V_{cb}V^*_{cd}+ V_{tb}V^*_{td} \approx V_{ub}+ \lambda V_{cb}+ V^*_{td} = 0$ The angles $\alpha$, $\beta$, and $\gamma$ are all, in principal, measurable from decays of $B$ mesons. The phenomonon of $B^0$ mixing is described in a similar way to the mixing of the neutral kaons: Initially pure $|B^0 \rangle$ and $|\overline{B}^0 \rangle$ states are written as orthogonal mixtures of heavy and light $B$ states. Because the heavy and light states each evolve with their own time- dependences, there is a quantum-mechanical oscillation of an initially pure $B^0$ or $\overline{B}^0$. The weak (CKM) phases in the mixing as well as in the decay amplitudes cause a CP asymmetry in the mixing and decay of initially pure neutral $B$'s as they time-evolve: \vspace{-0.1cm} \begin{equation} A_f(t) =\frac{\Gamma(B^0_{phys}(t) \rightarrow f) - \Gamma(\overline{B}^0_{phys}(t) \rightarrow f)} {\Gamma(B^0_{phys}(t) \rightarrow f) + \Gamma(\overline{B}^0_{phys}(t) \rightarrow f)} \end{equation} \vspace{-0.25cm} In the $B$ system it is possible to cleanly extract the weak phases from such asymmetries, in contrast with the situation in kaon mixing. In the Standard Model the asymmetry for $B_d^0$ mixing reduces to \cite{quinn,ali} \vspace{-0.25cm} \begin{equation} A_f(t) = - \mbox{Im} \alpha_f \sin(\Delta m_B t) \\ \end{equation} \vspace{-0.25cm} where \vspace{-0.25cm} \begin{equation} \alpha_f = \eta_{CP} \left(\frac{q}{p}\right)_B \frac{\langle f|H_{weak}|\overline{B}_q^0 \rangle}{\langle f|H_{weak} |B_q^0 \rangle} \left(\frac{q}{p}\right)_K \end{equation} \vspace{-0.05cm} and $\eta_{CP}=\pm 1$ is the CP sign of the final state $f$. The factor $(q/p)_B$ describes the weak phases in the $B$ mixing, and the factor $(q/p)_K$ appears whenever the final state $f$ has a $K_{\mbox{S}}$ or $K_{\mbox{L}}$ to account for kaon mixing in addition to $B$ mixing phases. The parameter $\Delta m_B$ is the $B_q^0-\overline{B}_q^0$ mass difference (for either neutral $B_s^0$ or $B_d^0$ mesons). For $B_d^0$ mesons, the phase $(q/p)_{B_d} = \mbox{arg}(V_{td}/V^{*}_{td})$ from the presence of intermediate top quarks in the box diagram describing $B_d-\overline{B}_d$ mixing.\cite{quinn,ali} The phase of the decay amplitudes depends upon the quark to which the $b$ decays. For $b\rightarrow c$ decays, there is no weak phase since $V_{cb}$ is almost real.\cite{frank2} Thus $b \rightarrow c$ decays in $B_d^0$ mixing gives us \vspace{-.25cm} \begin{center} $\alpha_f = \eta_{CP} \mbox{arg} \left( \frac{V_{td}}{V^*_{td}} \right)$, $A_f (t) = - \eta_{CP} \sin(2 \beta) \sin(\Delta m_B t)$ \end{center} \vspace{-.25cm} It should be noted that in extracting the value of $\sin (2\beta)$ from the mixing asymmetry, there is a four-fold ambiguity in the actual value of $\beta$ that is inferred from this measurement. In principle, present data, as shown in Figure 1, along with recent indications from the CDF experiment\cite{lewis}, indicate that $\mbox{sign}[\sin(2\beta)]>0$. It may be further possible\cite{babar} to furthermore remove the last two-fold ambiguity, as mentioned in Section VIII. In this paper I review several branching ratio measurements of $b \rightarrow c$ transitions of $B_d^0$ mesons that are relevant to future $B$ factories in measuring $\sin(2 \beta)$. The modes discussed are: $\begin{array}{lll} B^0 \rightarrow \psi K_{\mbox{S}} & B^0 \rightarrow \psi K_{\mbox{L}} & B^0 \rightarrow \psi(\mbox{2S}) K_{\mbox{S}} \\ B^0 \rightarrow \psi \eta & B^0 \rightarrow \chi_{\mbox{c}1} K_{\mbox{S}} & B^0 \rightarrow \psi \pi^0 \\ B^0 \rightarrow \psi K^{*0} & B^0 \rightarrow \psi(\mbox{2S}) K^{*0} & B^0 \rightarrow D^{*+}D^{*-} \\ \end{array}$ The above branching ratios were measured by the CLEO experiment running at the symmetric $e^{+}e^{-} \rightarrow \Upsilon\mbox{(4S)} \rightarrow B\overline{B}$ CESR collider at Cornell Unversity.\cite{urheim} The goal in studying these decays is to see how many events may be used for studying CP violation above and beyond the so-called ''gold-plated'' mode of $\psi K_{\mbox{S}}$. The last three decay modes are of course not CP eigenstates, but may still be used for studying CP violation, as is discussed below. The observation of CP violation in $B^0$ mixing and decay is a powerful first step toward proving CP violation is not simply a feature inherent to neutral kaons. The observation of this kind of CP violation will be an important confirmation of the CKM model\cite{frank}. \section{Data Sample} \vspace{-.15cm} Most of the studies presented in this paper represent the yield of 6.3 fb$^{-1}$ of data, which amounts to $\approx$ 6 million $B\overline{B}$ pairs. An additional 3 fb$^{-1}$ was taken 60 MeV below the $\Upsilon$(4S) to study backgrounds from continuum $e^{+}e^{-} \rightarrow q \overline{q}$ light quark production. At the $\Upsilon$(4S), it is convenient to use two kinematic constraints in reconstructing $B$ mesons from their daughter particles. Noting that the $B$ energy is exactly the $e^{\pm}$ beam energy, we form the invariant mass of $B$ candidates from $m_B^2 = \sqrt{E_{beam}^2 - (\Sigma {\bf p}_{i}^2)}$ since the beam energy is very well measured. Furthermore, we calculate the total energy $E_B$ of our $B$ candidates and form the variable $\Delta E = E_B-E_{beam}$ which should peak at zero for true signal. Resolutions on these quantities are $\sigma(m_B)\sim 2\mbox{--}3$ MeV and $\sigma(\Delta E)\sim 10\mbox{--}20$ MeV. While the analyses which identify final-state $\psi$ mesons differ slightly in their selection criteria, all of these studies were done by first selecting events where the $\psi$ decays to $ee$ or $\mu\mu$ pairs, which comprise just 12\% of the $\psi$ branching ratio. For these analyses, lepton selection criteria were imposed for both the daughter leptons in the $\psi$ decays, so the efficiencies for reconstructing the $\psi$ are typically $\sim 40\%$. In future $B$ factory experiments it may be possible to relax such selection criteria and thereby increase the reconstruction efficiencies for these decays.\cite{schuh} To suppress the $\psi$'s which come from continuum $e^{+}e^{-} \rightarrow c \overline{c}$, a cut of $P_{\psi} < 2.0 \mbox{GeV}/c$ was imposed. Thus, unless explictely mentioned otherwise, the backgrounds to the modes considered here from $B\rightarrow \psi X$ decays. CLEO has previously measured the inclusive $B\rightarrow \psi$ momentum distribution\cite{psipaper}, and our Monte Carlo model of $B$ decays, which includes several different exclusive $\psi$ modes, reproduces very well the observed $\psi$ momentum distribution. Therefore, many of the background estimates, which rely heavily on our Monte Carlo, are well modelled. \section{$B^0 \rightarrow \psi K_{\mbox{S}}$} \vspace{-.15cm} The decay $B^0 \rightarrow \psi K_{\mbox{S}}$ with $K_{\mbox{S}} \rightarrow \pi^+\pi^-$ is a so-called ''gold-plated'' CP eigenstate because the distinctive signature of two leptons from the $\psi$ and two charged tracks emanating from a point detached from the collision point. With efficiencies of $\sim 40\%$ for the $\psi$ and $\sim 75\%$ for reconstructing the $K_{\mbox{S}} \rightarrow \pi^+\pi^-$, we observe 75 events in 6.3 fb$^{-1}$, as shown in Figure 2(a). In Figure 2(a), we plot the $\Delta E$ of our $\psi K_{\mbox{S}}$ candidates {\it vs.} their mass. All quantities are plot in units of the experimental resolution (see Section 2), so signal is expected to lie in a region $\pm 3$ units from the expected values. Using our $B^0 \rightarrow \psi X$ MC, we expect $\leq 0.1$ events background in the sample. The branching ratio measured is $\mathcal{B}$$ (B^0 \rightarrow \psi K_{\mbox{S}}) = (4.6 \pm 0.06 \; \mbox{stat.} \pm 0.06 \; \mbox{sys.}) \times 10^{-4}$, where the first uncertainty is due to statistics and the second due to systematic uncertainties (in reconstruction efficiencies and the total number of $B\overline{B}$ in the sample from which these candidates were selected). It is hoped we can add to this 75 events using the other decay modes below. \section{$B^0 \rightarrow \psi K_{\mbox{S}}$, $K_{\mbox{S}} \rightarrow \pi^0\pi^0$} \vspace{-.15cm} To identify $K_{\mbox{S}} \rightarrow \pi^0\pi^0$ decays, we search for pairs of photons in the CsI calorimeter which are consistent with the $\pi^0$ mass using very mass cuts. When two such pairs are found, then the vector defined by the primary collision point and the center- of-energy of the four photons in the calorimeter is used to define the $K_{\mbox{S}}$ flight direction. The hypothesized $K_{\mbox{S}}$ flight distance before decaying is then varied until the two $\pi^0$ pairings give the best $\pi^0$ masses. The $K_{\mbox{S}}$ is not used in the constraint, but it is found that the $K_{\mbox{S}}$ mass resolution improves from $\sim 20$ MeV to $\sim 6$ MeV with this procedure. The $K_{\mbox{S}} \rightarrow \pi^0\pi^0$ reconstruction efficiency is 25\%, which, combined with the $\psi$ reconstruction efficiency, yields an overall efficiency of about 10\%. At an asymmetric $B$ factory one benefits from an additional constraint of matching the $K^0$ origin to the $B^0$ decay point as measured by the $\psi$ decay. We observe a signal of 15 events, with an expected background of just 0.7 events from what is believed to be random photons incorrectly paired with a $\psi$ from $B$ decays. However, we are investigating the possible contamination to the sample from $B^0 \rightarrow \psi K^{*}$ decays, since these can readily lend some photons and since the $\psi K^{*0}$ decay mode has a strong CP component. This yield, when combined with the reconstruction efficiency above, yields a branching ratio of $\mathcal{B}$$(B^0 \rightarrow \psi K_{\mbox{S}}) = (6.1 \pm 1.6 \; \mbox{stat.} \pm 0.13 \; \mbox{sys.}) \times 10^{-4}$ (see Table 1), consistent with our result in Section III. More importantly, this decay mode adds 15 more events to our sample of 75 for studying CP violation. \section{$B^0 \rightarrow \psi K_{\mbox{L}}$} \vspace{-.15cm} The $\psi K_{\mbox{L}}$ mode is interesting because in principle it presents the same (large) number of events for studying CP asymmetries as does $\psi K_{\mbox{S}}$, but the $\psi K_{\mbox{L}}$ has the opposite CP. It should therefore exhibit an asymmetry equal in magnitude but opposite in sign as the gold-plated mode. The difficulty is that the $K_{\mbox{L}} $ flight path is $\sim 1 \mbox{--} 2$ m, so it doesn't decay within the CLEO tracking volume. It is still plausible to detect the $K_{\mbox{L}}$'s, however, because the CLEO CsI calorimeter is 0.81 $\lambda_{\mbox{int}}$ in length, hence approximately 65\% of the $K_{\mbox{L}}$ interact in the calorimeter and initiate a shower that exceeds 100 MeV in energy. Thus, the signature for the $B^0 \rightarrow \psi K_{\mbox{L}}$ decay is the lepton pair from the $\psi$ plus a small calorimeter shower from the $K_{\mbox{L}}$. The $K_{\mbox{L}}$ shower is a nuclear interaction, is quite broad in comparison to showers from $\gamma$'s. In fact, often additional nearby showers are created as nuclear fragments travel some distance. Such shape distinctions are used in the selection to successfully reject over 90\% of $\gamma$ showers. In fact, of the 66 $\psi K_{\mbox{L}}$ candidates found, only 10 have showers that are due to random photons incorrectly paired with a $\psi$; all the rest are real $K_{\mbox{L}}$. The $K_{\mbox{L}}$ defines its direction, but not its energy. We use $E_{\psi}$ and ${\bf p}_{\psi}$ along with the $K_{\mbox{L}}$ direction to reconstruct $m_B$, but loose the $\Delta E$ constraint. To suppress backgrounds from $B^- \rightarrow \psi K^{(*)-}$ we reject events which have an additional track that makes a mass $\geq 5$ GeV. Furthermore, we veto events where the $K_{\mbox{L}}$ candidate shower makes the $\pi^0$ mass when paired with any other photon in the event in order to reject backgrounds from $B^0 \rightarrow \psi K_{\mbox{S}}$. Figure 2(b) shows the results of this search: 66 events are found in 6.7 fb$^{-1}$ of data, where 35 are expected to be from signal, 31 from backgrounds. As stated earlier, most of these backgrounds are from real $K_{\mbox{L}}$'s from $\chi_{\mbox{c1}} K_{\mbox{L}}$, $\psi K^{*0}$, and $\psi K^{*+}$. The CP dilution of these backgrounds is 15 events. We have not studied reconstruction efficiencies for this decay mode, although they can be inferred from the yield of 35 events in 6.6 fb$^{-1}$ and the known branching ratio for $\psi K_{\mbox{S}}$. \section{$B^0 \rightarrow \psi \pi^{0}$ / $\psi \eta$} \vspace{-.15cm} The $\psi \pi^{0}$ mode has the same CP and tests a similar Feynman diagram to the often-cited decay mode $B^0 \rightarrow D^{+}D^{-}$.\cite{babar} It is color-suppressed relative to $D^+D^-$, but perhaps $\psi \pi^0$ will yield more net events because the $D^+D^-$ channel has few subsequent decay modes which can be reconstructed. The $\psi \pi^0$ mode would presumably exhibit a asymmetry of $+ \sin(2\beta)$. Because it is the Cabibbo-suppressed version of the $\psi K^0$, we can predict the branching ratio will be $\mathcal{B}$$(B^0 \rightarrow \psi \pi^{0}) = (f_{\pi}/f_K)^2 \tan^2\theta_C$ $\mathcal{B}$$(B^0 \rightarrow \psi K^{0}) \sim 6 \times 10^{-5}$ or, by isospin conservation, $\mathcal{B}$$(\psi\pi^0) = 0.5\times$$\mathcal{B}$ $(\psi \pi^-) \sim (2.5\pm1.5)\times 10^{-5}$, where I've used the previously published CLEO result\cite{psipi} for $\psi \pi^-$. We reconstruct $\pi^0 \rightarrow \gamma \gamma$ decays in the CsI calorimeter, which has an efficiency of $\sim 60\%$. We observe 7 candidate events with the background expected to be 0.7 events. The background comes predominantly from real $\psi$'s from $B$ decay paired with random $\gamma$'s in the event which accidentally form the $\pi^0$ mass. With a net detection efficiency of 24.3\%, we obtain a branching ratio of $\mathcal{B}$$(B^0 \rightarrow \psi \pi^0) = (0.34 \pm 0.16 \; \mbox{stat.} \pm 0.04 \; \mbox{sys.})\times 10^{-4}$. We used the procedure of Feldman and Cousins \cite{feldman} to obtain the 68\% C.L. intervals for the Poisson signal mean. The $\psi \eta$ decay has the same CP as $\psi\pi^0$, so should exhibit the same asymmetry. By isospin, we might expect that $\mathcal{B}$$(\psi \eta) = \frac{1}{3} \times $$\mathcal{B}$$(\psi\pi^0)$. We searched for this mode using $\eta \rightarrow \gamma\gamma$ decays ($BR = 39\%$). In 6.3 fb$^{-1}$ no events were seen. \section{$B^0 \rightarrow \chi_{\mbox{c1}} K_{\mbox{S}}$} \vspace{-.15cm} This final state has the same CP as the $\psi K_{\mbox{S}}$, so should have the same sign asymmetry. We select this decay mode by searching for $\chi_{\mbox{c1}} \rightarrow \psi \gamma$, $\psi \rightarrow ll$, and $ K_{\mbox{S}} \rightarrow \pi^+\pi^-$ decays. In 6.3 fb$^{-1}$ 6 events were seen on a background of 0.6 events. The net detector efficiency is 16\%, giving a branching ratio of $\mathcal{B}$$(B^0 \rightarrow \chi_{\mbox{c1}} K^0) = (4.5^{+2.8}_{-1.8} \pm 0.9) \times 10^{-4}$. Again, the prescription of Feldman and Cousins \cite{feldman} was used. The background is expected to consist of random combinations of $\psi$'s and $\gamma$'s. We are investigating the explicit contribution from $\psi K^*$. \section{$B^0 \rightarrow \psi(2\mbox{S}) K^{(*)0}$} \vspace{-.15cm} The $\psi(2\mbox{S}) K_{\mbox{S}}$ mode is a CP eigenstate and the branching ratio is reported here for the first time. The $\psi(2\mbox{S}) K^{*0}$ mode has previously been observed by CDF\cite{cdfpsiprime} and is not a CP eigenstate: because the two vector particles originate from a spin-0 $B^0$, there is an additional factor of $(-1)^l = \pm 1$ in the final state CP due to orbital angular momentum between the particles. Even though the $\psi(2\mbox{S}) K^{*0}$ mode is a superposition of two CP eigenstates, it is hoped that a single CP state may dominate as with the decay $B^0 \rightarrow \psi K^{*0}$ earlier observed by CLEO and CDF.\cite{psikstar} Even if both CP states are prominent, however, Dunietz has suggested that an angular analysis may be used to separate the two CP components.\cite{dunietz} In this case, one must look at an angular distribution asymmetry that develops with proper decay time of the $B^0$ instead of just a decay rate asymmetry, so this analysis would require substantially more data. The $\psi$(2S) is reconstructed through its $l^+l^-$ ($BR = 12\%$) and $\psi \pi^+\pi^-$ ($BR = 32\%$) decays, and both $K^{*0} \rightarrow K^+\pi^-$ ($BR = 67\%$) and $K^{*0} \rightarrow K_{\mbox{S}}\pi^0$ ($BR = 17\%$) decays are considered. In this particular analysis the lepton identification was somewhat more stringent than the previous analyses, hence the efficiencies are somewhat lower (see Table 1). In the case of the $\psi(2\mbox{S}) K^{*0}$ mode, the systematic uncertainties are somewhat larger due to our knowledge of the unknown helicity amplitudes in this channel (we assume $\Gamma_{\mbox{L}}/\Gamma = 0.5$, and theoretically 0.5 - 0.7 is expected. Note that CDF and CLEO measure $\Gamma_{\mbox{L}}/\Gamma = 0.52 \pm 0.08$ for $\psi K^{*0}$ decays\cite{psikstar}). A total of 15 $\psi(2\mbox{S}) K_{\mbox{S}}$ and 21 $\psi(2\mbox{S}) K^{*0}$ candidates are observed (see Table 1). The backgrounds, totalling $0.4 \pm 0.2$ and $1.9 \pm 0.6$ events, respectively, are predominantly due to random combinatorics, but also from $B^- \rightarrow \psi(2\mbox{S}) K^{(*)-}$ decays. The relative yields of different $\psi$(2S) and $K^*$ decay modes is consistent with the different detection efficiencies and branching ratios. In addition to being a possible avenue for extracting $\sin(2\beta)$, the $\psi(\mbox{2S}) K^{*0}$ mode presented here and the $\psi K^{*0}$ earlier measured by CLEO\cite{psikstar} may help resolve two of the four-fold ambiguities in the extracted value of $\beta$. The time-development of $B^0\rightarrow \psi^{(')}K^{*0}$ decays contains additional terms due to the interference between the CP=+1 and CP=-1 components. This additional interference results in additional terms proportional to $cos(\beta)\sin(\Delta m_B t)$, hence a detailed measurement of this decay amplitude may eventually resolve the sign of $\cos(\beta)$ and remove two of the ambiguities for $\beta$. Unfortunately, the $K^{*0}\rightarrow K_{\mbox{S}}\pi^0$ decay mode has a quarter of the decay rate and less than half of the efficiency of the $K^+\pi^-$ mode, so gives a factor 9 fewer events in our sample; thus, such a resolution will not come concurrently with a measurement of $\sin(2 \beta)$. \section{$B^0 \rightarrow D^{*+}D^{*-}$} \vspace{-.15cm} Like the yet unobserved decay mode $B^0 \rightarrow D^+D^-$, this decay mode is not expected to be color-suppressed (as is $\psi \pi^0$), but it is expected to be Cabibbo-suppressed relative to the more prevalent and previously observed decay $B^0 \rightarrow D_{\mbox{s}}^{*+}D^{*- }$.\cite{dsdpaper} So, we might expect a branching ratio for $D^{*+}D^{*-}$ of approximately 0.1\%. Like the case of the $\psi^{(')}K^{*0}$, this mode is mixture of (+) and (-) CP eigenstates, so a mixing analysis works if one of the amplitudes dominates or one does a time-dependent measurement of an angular asymmetry\cite{dunietz}. A more complete description of this analysis has recently been published.\cite{jaffe} Here a brief description is given. To reconstruct this mode, both $D^{*+} \rightarrow D^0 \pi^+$ ($BR=67\%$) and $D^{*+} \rightarrow D^+ \pi^0$ ($BR = 33\%$) are used, although for background reasons only $B^0 \rightarrow (D^0\pi^+)(\overline{D}^0\pi^- )$ and $B^0 \rightarrow (D^+\pi^-)(\overline{D}^0\pi^-)$ modes were considered. A total of five $D^0$ and six $D^+$ decay modes were considered. To suppress background, two techniques were employed. The first was kinematic, in which a $\chi^2$ variable was constructed to compare the reconstructed $D$ and $D^*$ masses for candidates within a given event with the known values.\cite{pdg} The combinations of particles within an event with the best $\chi^2$ was chosen, and this combination had to have a $\chi^2<20$ (xx\% efficient for signal while xx\% efficient for random $B\overline{B}$ and $c\overline{c}$ backgrounds). The second technique was topological, requiring the flight distance between the $D$ and $\overline{D}$ decay vertices, as measured by the silicon vertex detector, to be inconsistent with zero given our experimental resolutions.\cite{svx} This cut was used for events where there was a slow $\pi^0$ from the $D^*$ decay. The results of the search are shown in Figure 3: 4 candidates were found. The background, expected to be 0.4 events, is due to $e^+e^- \rightarrow c\overline{c}$ continuum and to $B\rightarrow D+X$ decays, where a $D$ and a random $\pi$ are incorrectly paired to make a $D^*$. The probability of 0.4 events expected background fluctuating to 4 observed events is $1.1 \times 10^{-4}$. The branching ratio is determined to be $(6.2^{+4.0}_{-2.9}\pm1.0) \times 10^{-4}$, consistent with our expectations and actually quite comparable to the $\psi K_{\mbox{S}}$ mode. Unfortunately, the reconstruction efficiency is quite small here, just $\sim 10^{-3}$, which is in part due to the 8\% from all the branching ratios of the $D$'s and $D^*$'s, but also to the 1\% efficiency for reconstructing these high multiplicity decays. We are currently investigating the possibility of only partially reconstructing one of the $D^*$'s in this decay, and it is furthermore possible that the reconstruction efficiency will be higher at an asymmetric $B$ factory due to the Lorentz boost of the $B$. \section{Summary} \vspace{-.15cm} Table 1 shows the yields for the CP eigenstates summarized in this paper. In 6.6 fb$^{-1}$, 140 events are found amongst all the modes after dilutions are accounted for. Thus, in a 30 fb$^{-1}$ dataset which might be accumulated by one of the $B$ factories in one year's run, one could anticipate 640 events for studying $\sin (2\beta)$. Furthermore, as discussed in Section 2, some of the selection criteria in these various searches have not been optimized for CP studies, hence the efficiencies might be increased by another factor of $\sim 1.6$, giving an anticipated number of events (after dilution) to just over 1000 when all modes are included. Furthermore, one can hope for additional background rejection when working at an asymmetric collider. Even ignoring such gains, however, one should be able to measure $\sin(2\beta)$ with an error of $\pm 0.1$. The fact that the 3-generation Standard Model with a single Higgs multiplet has just one independent CP-violating phase makes all the CP-violating effects all very strongly constrained. It will be of great interest to see whether the pattern of CP violation in $B$ decays agrees with the prediction of the CKM model, or whether new physics will have to be invoked to understand the (hopefully many!) manifestations of CP violation observed in the next few years. \section{Acknowledgements} \vspace{-.15cm} It is a pleasure to thank my many CLEO colleagues for the opportunity to represent them at the Division of Particles and Fields conference. The January Los Angeles climate is a pleasant departure from Upstate New York. In particular, I thank Alexey Ershov, Alex Undrus, Andy Foland, and David Jaffe, whose work is represented here. I also thank Sheldon Stone and Silvia Schuh for helpful discussions.
\section{Introduction} Recent advances in infrared and X-ray array technologies have greatly increased the spatial coverage and resolutions in near-IR and X-ray observations, providing powerful tools to compare high-resolution radio, IR and X-ray images of the complex region of the Galactic center. Sgr A$^*$ is a bright compact radio source which is considered to be the strongest candidate for a 2.5$\times10^6$ \msol black hole at the dynamical center of the Galaxy (Eckart and Genzel 1997; Ghez et al. 1998). This source is surrounded by diffuse ionized gas as well as clusters of hot and cool stars seen in near-IR wavelengths. Detailed comparison of radio and near-IR images was recently made with a new approach by Menten et al. (1997), who detected a number of SiO (0.7 cm) and H$_2$O (1.3 cm) masers associated with known compact stellar sources seen in the diffraction limited 2.2 $\mu$m infrared images. Astrometric measurements between radio and 2 $\mu$m images of this region were used to determine the position of Sgr A$^*$ with respect to other well-known near-IR sources. Although the relative position of Sgr A$^*$ is accurately measured within 30 mas in this technique, the absolute position of Sgr A$^*$ was known with an accuracy of 0.2$''$. A more accurate position of this source and its subsequent X-ray counterpart in future sub-arcsecond observations has important implications on the nature of high energy activity predicted from models of this source (Melia 1994; Narayan et al. 1995). The absolute positions of radio sources are routinely obtained to an accuracy of a few milliarcseconds using centimeter wavelength VLBI techniques. However, the large scattering diameter of Sgr A$^*$ (Yusef-Zadeh et al. 1994) makes the application of centimeter wavelength VLBI less feasible, as the source is completely resolved on longer baselines. Thus, shorter baseline arrays and/or shorter wavelengths are needed resulting in reduced positional accuracy. The most recent measurements of the location of Sgr A$^*$ with the smallest error bar is given as $\alpha$, $\delta$[ 1950 ] = $17^{\rm h} 42^{\rm m}$ 29\dsec314$\pm0.010$, $-28^\circ 59^\prime 18.3\pm0.2^{\prime\prime}$ (Rogers et al. 1994). This position is determined with respect to the reference source NRAO 530 using VLBI techniques at 3mm. To improve the absolute position of Sgr A$^*$, we obtained archival VLA data sets at a number of frequencies taken over the last 13 years. By averaging the position of Sgr A$^*$ at different frequencies and at different epochs, the new position of Sgr A$^*$ is determined with an accuracy of an order of magnitude better than earlier measurements. The averaging of the data over such a long span of time and frequency can be justified. The proper motion measurements of this source is too small with respect to reference extragalactic radio sources (Reid et al. 1998; Backer 1996) to affect individual measurement. Also, the change in the position of Sgr A$^*$ as a function of frequency is too small to be significant compared to other systematic errors. The frequency dependence of the size of Sgr A$^*$ due to scattering of radio waves will not change the centroid of the position of Sgr A$^*$ as a function of frequency (Yusef-Zadeh et al. 1994). It is worth mentioning that the detection of frequency dependence of the position of Sgr A$^*$ at a milliarcsecond level, due to its intrinsic radio emission, has the potential to support the asymmetrical Bondi-Hoyle accretion model (Melia 1994). \section{Observations} Table 1 shows the list of 15 data sets that were based on continuum observations made with the Very Large Array of the National Radio Astronomy Observatory\footnote{The National Radio Astronomy Observatory is a facility of the National Science Foundation, operated under a cooperative agreement by Associated Universities, Inc.} at 6 (5 GHz), 3.6 (8.4 GHz), 2 (15 GHz), 1.2 (20 GHz), and 0.7 cm (44 GHz) between 1986 and 1999. All had a bandwidth ranging between 2 $\times$ 50 and 2 $\times$ 12.5 MHz in each right and left circular polarization. Each of these original data sets had relatively good {uv} coverage and none were done in snapshot mode with less than one hour of integration time on the source. With the exception of 0.7 cm data which was taken in the C configuration, all the observations were done in either the A or the B configurations of the VLA to obtain the highest resolution images of Sgr A$^*$ with minimal contamination or confusion by ionized features in the vicinity of Sgr A$^*$. The low resolution images of Sgr A$^*$ at 20 and 90 cm suffer from this source confusion as well as optical depth effects due to intervening thermal and nonthermal continuum sources. Thus, 20 and 90cm observations were not used. In all observations, NRAO530 and 3C286 were chosen as the phase (astrometric) and flux (photometric) calibrators, respectively. All processing was done using the NRAO AIPS analysis package; the standard calibration was applied and the field around Sgr A$^*$ was imaged. Since the observations were all made in the B1950 coordinate system, the data were analyzed in this system. The VLA calibrator B1950 coordinate system is referred to epoch 1979.5. No self--calibration was done on any of the data as this procedure loses the absolute position of a source in order to minimize atmospheric phase errors. Prior to self-calibration, no significant emission from Sgr A$^*$ was detected at 0.7 cm in the A and B configurations, so these data were excluded. VLA continuum data consists of two independent frequency bands. The final images based on combining these two frequency bands were used to determine the position of Sgr A$^*$ with AIPS routine JMFIT by fitting a Gaussian to the source. Fits done using images from the individual frequency bands showed no significant positional shift. At short wavelengths, the phase fluctuations degrading the positional accuracy are due mainly to variable refraction in the troposphere; at longer wavelengths the ionosphere adds increasingly important contributions. Variable ionospheric refraction is particularly significant for observations made at 6 cm during the day when the ionosphere is densest. Solar heating of the troposphere also increases the variations in refraction. So, we have assigned weights of 1 and 2 to day-- and night-time observations, respectively. For observations that were carried out between night and day, the assigned weight was 1.5. The atmospheric effects are aggravated by the low declination of the source and the calibrator and the resultant low elevation angles of the observations. These effects are minimized by restricting the {\it uv} data to near transit. The data sets selected within one to three hours of the highest elevation improved considerably the dispersion in measured position of Sgr A$^*$ among the different epochs. Errors in atmospheric refraction affect mainly the measurements of the declination of Sgr A$^*$. The astrometric calibrator used, NRAO530, is about 15$^\circ$ north of Sgr A$^*$. Since the observations were restricted to those near transit, this corresponds to a difference primarily in elevation. Corrections to the atmospheric refraction model are determined from the calibrator observations which is at a significantly higher elevation than Sgr A$^*$ (i.e. at lower air mass); any unmodeled differences in atmospheric refraction between the calibrator and Sgr A$^*$ will appear primarily as errors in declination. A further potential source of systematic error is the assumed position of the calibrator, NRAO530. Our measurements are of position relative to that of the calibrator whose absolute position is known; any errors in the assumed position of NRAO530 will directly affect the derived position of Sgr A$^*$. In all our measurements, the position of NRAO530 was constant and was assumed to be at (B1950 epoch 1979.5) $\alpha$, $\delta$ = $17^{\rm h} 30^{\rm m} 13\dsec5352$, $-13^\circ 02^\prime 45.837^{\prime\prime}$. This position is consistent to within 5 mas of the VLBI source position given by Johnston et al. (1995) and by US Naval Observatory's (USNO) astrometry list (ftp://casa.usno.navy.mil/navnet/n9798.sor) after the correction between the VLA and USNO frames is made. The position of NRAO530 assumed is an absolute position measured by VLBI techniques which are insensitive to structure larger than a few tens of milliarcseconds. NRAO530 is suspected of having structure on the scale of several hundred milliarcseconds which would not affect the VLBI measurements but will cause an apparent shift of the position when viewed with the lower VLA resolution. Any difference between the centroid of the position of NRAO530 at VLBI and VLA resolutions will appear as an error in the derived position of Sgr A$^*$. \section{Results} Table 1 lists 15 data sets, three of which at 1.3 cm, four at 2 cm, five at 3.6 cm, two at 6 cm, and one at 0.7 cm. The date of each observation, the Gaussian fitted position with 1 $\sigma$ error, duration of the observation and the corresponding weight in the final averaging are also given. The errors of the fitted positions depend on the signal-to-noise ratio and are much less than the errors introduced by the atmosphere. The rms scatter within the data set, as described below, is about ten times the fitted noise, thus we have not used the errors of the fitted positions in any of our analysis. The mean position of Sgr A$^*$, both weighted and unweighted, as well as the position determined by Rogers et al. 1994 are also listed at the bottom of the table. The most descripant point at 1.3 cm was not included in the final averaging. The rms scatter of the individual measurements in Table 1 (either weighted or unweighted) is about 0.035$''$ in Right ascension and 0.055" in declination. This is consistent with the assumption that the errors are dominated by uncertainties in atmospheric refraction which should produce a larger scatter in declination. This is further supported by the two pairs of positions measured on the same day, in which the declinations differed by about 0.030", or nearly half of the day-to-day scatter. The one sigma errors in the mean position are taken to be the rms scatter of the sample divided by the square root of the number of observations minus one. Figure 1 shows the Right ascension and declination of Sgr A$^*$ with one sigma error given by Rogers et al. (1994) against 15 positions determined from our measurements. All the new positions fall to the SW quadrant of the position derived by Rogers et al. The weighted position of Sgr A$^*$ with the weighted mean epoch of our observation being 1992.4 is \hfill\break (B1950) $\alpha$, $\delta$ = $17^{\rm h} 42^{\rm m}$ 29\dsec3076$\pm0.0007$, $-28^\circ 59^\prime 18.484\pm0.014^{\prime\prime}$ or \hfill\break (J2000) $\alpha$, $\delta$ = $17^{\rm h} 45^{\rm m}$ 40\dsec0383$\pm0.0007$, $-29^\circ 00^\prime 28.069\pm0.014^{\prime\prime}$. \hfill\break These measurements are in good agreement with the less precise measurement of Rogers et al. 1994 as well as with the position of Sgr A$^*$ determined from recent VLBI measurements (Reid et al. 1999).
\section{Introduction} Toy models in physics play an important role in understanding the basic features of more involved theories and phenomena. In particular, models in one dimensional quantum mechanics, illustrating analogous situations in a field theoretic context have been quite useful for advanced researchers as well as beginning graduate students. Among many such models, those which illustrate non--perturbative aspects of field theory through an analysis of instanton solutions and bounces have been looked at in diverse contexts. The $x^{2}-x^{4}$ potential and the quantum pendulum have been discussed and analyzed in great detail in the past. Apart from aspects which emerge out of solutions and their analysis, it is also important to relate the toy model with different realistic models by mapping rules/dimensional extensions etc. We shall, in this article, try to analyze a novel toy model which has applications in different situations and is also exactly solvable to some extent. In particular, we discuss the problem of {\em Particle on a Rotating Circle} (henceforth referred to as PORC) in the presence of gravitational/ magnetic field (a charged particle in the case of magnetic field). Before we indulge into analyzing the salient features of this model, let us first explore the existing literature on it. For a undergraduate/graduate student, a first encounter with this model is likely to occur while doing a course in classical mechanics. The model appears as a problem in the second chapter of Goldstein's book on classical mechanics{\cite{gold:cm}}. It is also elaborately discussed in Arnold {\cite{gold:cm}}. There exists a host of other articles on it, in this journal too {\cite{porc:ajp}}, the most recent one being published a couple of years ago {\cite{flet:ajp97}}. The major directions along which this model has been viewed are (i) as an example of spontaneous symmetry breaking (ii) as a mechanical model for second order phase transitions (iii) as an example in quantum mechanics where instanton solutions are obtainable {\cite{sk:pla92}}(iv) as a lower dimensional analog of certain higher dimensional field theories. In this article, we shall see that this model (PORC) has the following intriguing features : \begin{itemize} \item{The classical mechanics of the model is exactly solvable in a way similar to that of the pendulum problem } \item{When extended to full real line (instead of circle $S^{1}$)\ the model has intimate connections with several field theoretic and statistical mechanical models (for details see Section III).} \item{The model in the presence of a uniform gravitational or magnetic field has non-perturbative solutions such as instantons/bounces and the corresponding tunneling/decay rates can be written down and analyzed (for a discussion on the theory of instantons, bounces and their contribution to tunneling/decay rates can be found in \cite{gen:inst}, \cite {gen1:inst}, \cite{kleinert:pi} ).} \item {A formal analysis of the Schr\"odinger equation can be performed and exact solutions obtained. However the full expressions for the eigenvalues and eigenfunctions are encoded in certain continued fractions which we do not discuss in further detail.} \item{In the presence of a uniform magnetic field the quantum problem is quasi-exactly solvable--we can get at least a few energy eigenstates and eigenvalues analytically \cite{hks} \cite{km} by extending the method due to Razavy {\cite{raz:ajp}}.} \end{itemize} \section{The model} Let us first introduce the model in some detail. We have a point particle of mass $m$ (and charge $q$ if we are looking at the problem in the presence of a magnetic field) constrained to move on a circle ($S^{1}$). The circle is located vertically and is rotating at an angular frequency $\omega$ about the vertical axis of symmetry. Two special cases will be analyzed -- (i) the system in the presence of a uniform gravitational field (ii) the system in the presence of a uniform magnetic field. Specifications about the directions and magnitudes of each of these uniform fields are given below. Note that this problem is a generalisation ($\omega \neq 0$) of the usual pendulum problem ($\omega =0$) whose quantum mechanics is discussed in {\cite{kp:ajp}} A generic Lagrangian for the such a system can be given as : \begin{equation}\label{2.1} L_{generic} = \frac{1}{2}mr^{2}{\dot \theta}^{2} - A\cos \theta - B\cos 2\theta \, , \end{equation} where A and B are constants depending on the various parameters in the model. We can in principle have four cases depending upon the signs of A and B. These are : \centerline{(i) $A>0, B>0 $ (ii) $A>0, B<0$ (iii) $A<0,B>0$ (iv) $A<0, B<0$} Note that with a proper redefinition of $\theta$ (more precisely $\theta \rightarrow \pi - \theta$) we can relate the models (i) and (iii) as well as (ii) and (iv). Therefore, in essence we have only two models to discuss. The two special cases given below exemplify these two cases. {\bf Case 1 : System in a uniform gravitational field} \begin{equation} L_{grav} = \frac{1}{2}mr^{2}\dot \theta^{2} - \bigg (mgr \cos \theta - \frac{1}{2} mr^{2} \omega^{2}\sin^{2}\theta \bigg ) \, , \end{equation} where $\theta$ is the generalized coordinate required to describe the system. Thus the effective one--dimensional potential is : \begin{equation} V(\theta) = mgr\cos\theta +\frac{1}{4}mr^{2}\omega^{2}\cos 2\theta - \frac{1}{4} mr^{2}\omega^{2} \, . \end{equation} Note that this system falls in the $A>0,B>0$ class mentioned above. The location of the extrema of the effective potential are as follows : {\bf Minima :} (i) $\omega > \omega_{0}$ : $\theta = \cos^{-1}{(-a)} \, , 2\pi - \cos^{-1}(-a) \, .$ $V(\theta_{min}) = V(2\pi - \theta_{min}) = -mr^2 \omega_0^2 (a^2+1)/a \, .$ (ii) $\omega <\omega_{0}$ : $\theta =\pi$ \hspace{.1in} ; \hspace{.1in} $V(\theta_{min}) = -mr^2 \omega_0^2 \, .$ {\bf Maxima :} (i) $\omega >\omega_{0}$ : $\theta = 0(2\pi)$, $\theta =\pi$ \hspace{.1in} ; \hspace{.1in} $V(0) = V(2\pi) = mr^2 \omega_0^2 \, ; \ V(\pi) = -mr^2\omega_0^2 \, .$ (ii) $\omega <\omega_{0}$ : $\theta = 0 (2\pi) \, .$ Thus for $\omega > \omega_0$ one has degenerate minima and maxima as well as a local maxima while for $\omega < \omega_0$ there is no local maxima. The potentials are shown in Fig. 1(a) and 1(b) for the $\omega>\omega_{0}$ and $\omega<\omega_{0}$ cases respectively. Here $\omega_0 = \sqrt{g/r}$ while $a = \omega_0^2/\omega^2$. {\bf Case 2 : System in a uniform magnetic field} \begin{equation} L_{mag} = \frac{1}{2}mr^{2}{\dot\theta}^{2} + \frac{1}{2}mr^{2}\omega^{2} \sin^{2}\theta + q {\bf A . v} \, , \end{equation} where $q$ denotes the charge of the particle. Since $\bf B$ is constant we have ${\bf A}= -\frac{1}{2} {\bf r \times B}$. Assuming that the components of $\bf B$ are $(B_{r}, B_{\theta},0)$ we find that the vector potential $\bf A$ has the components $(0,0,A_{\phi})$ where $A_{\phi} = -\frac{1}{2} r B_{\theta}$. Since ${\bf v}$ has components $(0,r\dot\theta, \omega r \sin \theta)$ we find that the Lagrangian takes the form : \begin{equation} L_{mag} = \frac{1}{2}mr^{2}{\dot \theta}^{2} - \left ( \frac{qBr^{2}\omega}{2} \sin \theta - \frac{1}{2} mr^{2}\omega^{2} \sin^{2} \theta \right ) \, . \end{equation} Thus the effective one dimensional potential is \begin{equation} V (\theta ) = \frac{1}{2}mr^{2}\omega^{2} \left ( \frac{\omega_{c}}{\omega} \sin \theta -\sin^{2}\theta \right ) \, , \end{equation} where $\omega_{c} = \frac{qB}{m}$. With a straightforward redefinition of $\theta$ ($\theta \rightarrow \theta - \frac{\pi}{2}$) we see that this model belongs to the other class ($A>0,B<0$). The extrema of the effective potential are as follows : {\bf Minima :} (i) $\omega > \omega_c/2 : \theta = \frac{\pi}{2} \, , \frac{3\pi}{2} \, ,$ $V (\frac{\pi}{2}) = \frac{1}{2}mr^{2}\omega^{2} \left ( \frac{\omega_{c}}{\omega} - 1 \right ) \quad ; \quad V (\frac{3\pi}{2}) = -\frac{1}{2}mr^{2}\omega^{2} \left ( \frac{\omega_{c}}{\omega} + 1 \right )$ (ii) $\omega < \omega_c/2 : \theta = {3\pi\over 2} \quad ; \quad V(3\pi/2) = - {1\over 2} mr^2 \omega^2 ({\omega_c \over \omega} +1)$ {\bf Maxima :} (i) $\omega > \omega_c/2 : \theta = \sin^{-1}\frac{\omega_{c}}{2\omega}, \ \pi - \sin^{-1}\frac{\omega_{c} }{2\omega} \quad ; \quad V (\theta_{max}) = \frac{1}{8}mr^{2} \omega_{c} ^{2}$ \, . (ii) $\omega < \omega_c/2 : \theta = {\pi \over 2} \quad ; \quad V(\pi/2) = {1\over 2} mr^2 \omega^2 ({\omega_c \over \omega} -1)$ \, . Thus for $\omega > \omega_c /2$ the effective potential represents a system which has a false vacuum at $\theta = \frac{\pi}{2}$ and a true vacuum at $\theta = \frac{3\pi}{2}$ while for $\omega < \omega_c /2$ there is no local minima. The two scenarios are plotted in Fig. 2(a) and 2(b) respectively. >From the figures for the effective potentials we can conclude the following -- (i) in the presence of a uniform magnetic field we may have a true as well as false vacua while (ii) in a uniform gravitational field we can have a degenerate double well potential. Additionally, if we generalize the above over the full (real) line then we have a periodic potential for which the minima are infinite-fold degenerate. A more general problem is that of a charged particle in a uniform gravitational plus magnetic (and even electric) field. The most general Lagrangian is given by \begin{equation}\label{2.7} L_{general} = \frac{1}{2}mr^{2}{\dot \theta}^{2} +\frac{1}{2}mr^{2} \omega^{2} \sin^{2} \theta - mgr\cos\theta -\frac{qBr^{2}\omega}{2} \sin \theta + q\phi_{elec} \, , \end{equation} where $\phi_{elec}$ is the electric potential. In general, this problem cannot be exactly solved in either classical or quantum mechanics even though a general analysis of the motion in classical mechanics is possible {\cite{flet:ajp97}}. However, by suitably choosing the ratio of the electric and magnetic fields we can reduce the problem to that of a particle moving on a rotating circle in uniform gravitational field alone. Similarly, by suitably choosing the ratio of the gravitational and electric fields, one can reduce it to that of a particle moving on a rotating circle in a uniform magnetic field alone. Other combinations and their consequences can also be tried out. We leave it to the reader to figure out the details of these scenarios. \section{Inter--connections with field theoretic and other models} The toy models discussed above have remarkable connections with a wide--ranging variety of field theoretic models. We shall now briefly summarize some of these interconnections. To begin with, note that there exists a correspondence between the toy models discussed above and $1+1$ dimensional double-- sine--Gordon (DSG) field theory defined by the Lagrangian density : \begin{equation}\label{3.1} {\cal L} = {1\over 2}{\partial_{\mu} \phi} {\partial^{\nu}\phi} + Acos \phi + B cos 2\phi \, , \end{equation} in that the potential is formally the same in both the cases. Here A and B could be positive or negative. The field equation that follows from a first variation is \begin{equation}\label{3.2} {\partial^2 \phi \over \partial t^2} - {\partial^2 \phi \over \partial x^2} = A \sin \phi + 2B \sin 2\phi \, . \end{equation} Hence the equation for static (i.e. $\phi$ independent of time) solutions is \begin{equation}\label{3.3} {d^2 \phi \over dx^2} = -\left ( A sin \theta + 2B sin 2\theta \right ) \, . \end{equation} It may be noted that this equation also follows from our toy model Lagrangian as given by eq. (\ref{2.1}) but for an overall sign. In fact if we go to the Euclidean time i.e. let $t = i\tau$ , then the classical equation of motion in the toy model takes the form \begin{equation}\label{3.4} {d^2 \theta \over d\tau^2} = - (A sin \theta + 2B sin \theta) \, , \end{equation} which is identical to the static field equation as given by eq. (\ref{3.3}). Now it is well known {\cite{gen1:inst}} that the finite Euclidean action solutions of the Euclidean equations of motion (\ref{3.4}) for any system are nothing but the instantons, whose mere existence is due to the presence of degenerate minima in the potential appearing in the theory. Further, once one has obtained these instanton solutions and the corresponding Euclidean actions, the tunneling amplitude for the transition between the degenerate vacua can easily be computed in the dilute (non-interacting) instanton gas approximation. We thus notice that the static finite energy solutions of the DSG equation as given by (\ref{3.3}) are the same as the finite (Euclidean) action instanton solutions of our Toy model (with the obvious replacement of $\phi$ with $\theta$ and $x$ with $\tau$). Since the DSG equation has been extensively studied in the literature {\cite{leu,ptk,boc,cgm,det}}, we can immediately write down the corresponding instanton solutions of our toy model. It may be added here that the DSG field equation occurs quite naturally in several physically important situations. For example, it appears quite naturally in the study of spin waves in the B-phase of super-fluid $^3 He$ \cite{mat,buc}, in the problem of self-induced transparency \cite{dbc} and in nonlinear excitations in a compressible chain of XY dipoles under conditions of piezoelectric coupling \cite{rem}. Further, a model based on the DSG theory has been proposed \cite{hls} to describe the poling process in the B-phase of the polymer polyvinylidene fluoride ($PVF_2$). Additionally, the DSG theory has been investigated as a model of a nonlinear system which can support more than one kind of soliton. The sine-Gordon equation, in turn, is related to the massive Thirring Model \cite{col} while a specific form of DSG has been shown to be equivalent to a generalized massive Thirring model \cite{bha}. Thus, any new results obtained in any of these models will immediately have relevance in various other related models. Finally, we may add that the DSG equation is also related to the anisotropic Heisenberg chain in an applied magnetic field. In particular, by treating spins classically, Leung \cite{leu} as well as Pandit et al. \cite{ptk} have shown that the classical spin dynamics is approximately governed by the generalized double sine Gordon (DSG) model as given by eq. (\ref{2.7}). \section{Classical Solutions} We now move on towards analyzing the real time solutions i.e. the classical mechanics of the above problem. Defining $E=T+V-V_{min}$ so that $E \ge 0$, we integrate the corresponding quadrature and then analyze the various cases separately. {\bf 1. System in a gravitational field} For the gravitational case, for the two domains of $\omega$ we have the following results.. {\underline{\sf $\omega <\omega_{0}$}} Here $V_{min}=-mr^{2}\omega_{0}^{2}$. Using $\theta = 2\phi$ and then $\tan \phi = y$ we obtain the following generic form of integral which we need to analyze : \begin{equation} \int \frac{dy}{\sqrt{y^{4} + 2y^{2}\left (1-\alpha -\beta \right ) + \left (1-2\alpha\right )}} = \sqrt{\frac{E}{2mr^{2}}} \left ( t-t_{0}\right ) \end{equation} where \begin{equation} \alpha = {m\omega_0^2 r^2\over E} \quad ; \quad \beta = {m\omega^2 r^2 \over E} \, . \end{equation} Recall that $V_{max}=2mgr=2mr^{2}\omega_{0}^{2}$. We can therefore have two possibilities $E<V_{max}$ and $E>V_{max}$ which we discuss one by one. {\sf (a) $E<V_{max}$} In this energy range we have the following two alternatives depending on the sign of the quadratic term in the denominator of the elliptic integral. These are : (i) $0<E<mr^{2}\left (\omega_{0}^{2}-\omega^{2} \right )$ In this case the solution is \cite{gr} \begin{equation} \tan \frac{\theta}{2} = b \frac{dn(X,k)}{\kappa^{\prime}sn(X,k)} \end{equation} where $X=\frac{b}{\kappa^{\prime}}\sqrt{\frac{E}{2mr^{2}}}(t-t_{0})$, $a^{2}b^{2}=2\alpha -1$, $ b^{2}-a^{2} = 2 (\alpha -\beta -1)$ \, . Here cn(X,k), sn(X,k) and dn(X,k) are Jacobi elliptic functions of real elliptic modulus parameter $k$. (ii) $mr^{2}\left (\omega_{0}^{2}-\omega^{2} \right ) < E <V_{max}$ In this case the solution is \begin{equation} \tan {\frac{\theta}{2}} = \frac{a dn(X,k)}{\kappa sn(X,k)} \end{equation} where $X=\frac{a}{\kappa}\sqrt{\frac{E}{2mr^{2}}}(t-t_{0})$, $ \frac{a}{a^{2}+b^{2}} =\kappa$, $\kappa^{\prime} = \sqrt{1-\kappa^{2}}$, $a^{2}b^{2} = 2\alpha -1$, $a^{2}-b^{2}=2(1-\alpha+\beta)$. {\sf (b) $E>V_{max}$} The exact solution here is : \begin{equation} \tan \frac{\theta}{2} = a cn(X,k)/sn(X,k) \end{equation} where $X= a\sqrt{\frac{E}{2mr^{2}}}(t-t_{0})$, $\frac{\sqrt{a^{2}-b^{2}}}{a} =\kappa$, $a^{2}b^{2}=1-2\alpha$, $a^{2}+b^{2}=2(1-\alpha+\beta)$. Note also that as $b\rightarrow 0$, $a^{2}=1+\frac{\omega^{2}}{\omega_{0}^{2}}$ , which implies $E=2mr^{2}\omega_{0}^{2}$. {\underline{\sf $\omega >\omega_{0}$}} In this domain of $\omega$ we note that $V_{min}= -\frac{mr^{2}}{2\omega_{0}^{2}} \left (\omega_{0}^{4} +\omega^{4} \right ) $. As before, using $\theta = 2\phi$, $\tan \phi = y$ we get : \begin{eqnarray} \int \frac{dy}{\sqrt{y^{4} \left ( E-V_{max} + 2mr^{2}\omega_{0}^{2} \right ) + 2y^{2} \left ( E + mr^{2}\omega_{0}^{2} + mr^{2}\omega^{2} -V_{max} \right ) + \left ( E-V_{max} \right ) }} \nonumber \\ = \sqrt{\frac{1}{2mr^{2}}} \left (t-t_{0} \right ) \, , \end{eqnarray} where $V_{max} = \frac{m r^2}{2 \omega^2} \left (\omega_{0}^{2} + \omega^{2} \right )^{2}$. The exact solutions in the various sub-cases are analyzed below. {\sf (a) $E<V_{max}$} In a way similar to the $\omega<\omega_{0}$ case we once again have two possibilities depending on the signs of the coefficients of the various terms appearing in the elliptic integral. (i) $0<E<\frac{mr^{2}}{2\omega^{2}}\left (\omega^{2}-\omega_{0}^{2} \right )^{2}$ The solution for this case is : \begin{equation} \tan \frac{\theta}{2} = \frac{1}{a dn(X,k)} \end{equation} where $X= a\sqrt{\frac{V_{max}-E}{2mr^{2}}} (t-t_{0})$, $a^{2}b^{2} = \frac{V_{max} - E - 2mr^{2}\omega_{0}^{2}}{V_{max} - E}$, $a^{2} + b^{2} = 2\frac{E-V_{max}+mr^{2}(\omega^{2}+\omega_{0}^{2})} {V_{max} - E}$. (ii) $\frac{mr^{2}}{2\omega^{2}}\left (\omega^{2}-\omega_{0}^{2} \right )^{2}< E <V_{max}$ Within these bounds of energy, the solution turns out to be \begin{equation} \tan {\frac{\theta}{2}}= \frac{a}{\kappa} \frac{dn(X,k)}{sn(X,k)} \end{equation} with $X= \frac{a}{\kappa} \sqrt{\frac{E+2mr^{2}\omega_{0}^{2} - V_{max}} {2mr^{2}}} (t-t_{0})$, $a^{2}b^{2} = \frac{V_{max}-E}{E+2mr^{2}\omega_{0}^{2} -V_{max}}$, $a^{2}-b^{2} = \frac{E+mr^{2}(\omega^{2}+\omega_{0}^{2})-V_{max}} {E+2mr^{2}\omega_{0}^{2} -V_{max}}$. {\sf (b) $E>V_{max}$} The solution here is given by: \begin{equation} \tan{\frac{\theta}{2}} = a\frac{cn(X,k)}{sn(X,k)} \end{equation} where $X=a\sqrt{\frac{E-V_{max} +2mr^{2}\omega_{0}^{2}}{2mr^{2}}} (t-t_{0})$. {\bf 2. System in a magnetic field :} In the magnetic case, we follow the same procedure as above. The only major difference is that, unlike the gravitational case, $V_{min} = -\frac{1}{2} mr^{2}\omega^{2} \left ( 1 +\frac{\omega_{c}}{\omega} \right )$ is the same irrespective of whether $\omega>\frac{\omega_{c}}{2}$ or $\omega < \frac{\omega_{c}}{2}$. Using $\theta = \phi -\frac{\pi}{2}$, $\frac{\phi}{2} = \eta$ and $\tan \eta =y$ we find that we have to handle the following integral : \begin{equation} \int \frac{dy}{\sqrt{\left [ \left (1-\alpha\right ) y^{4} + \left (2-\alpha -\beta \right ) y^{2} + 1 \right ]}} = (t-t_0) \sqrt{E\over 2mr^2} \, , \end{equation} where $\alpha = {mr^2 \omega \omega_c \over E}, \beta = {2mr^2 \omega^2 \over E}$. It is convenient to make another transformation $y=\frac{1}{u}$ and then analyse the resulting integral, which is, generically, of the form \begin{equation} -\int \frac{du}{\left [ u^{4} + u^{2} \left ( 2-\alpha -\beta \right) + \left ( 1-\alpha \right ) \right ]^{\frac{1}{2}}} = \left (t-t_{0}\right ) \sqrt{\frac{E}{2mr^{2}}} \, . \end{equation} The various cases are now analyzed below. {\underline{\sf $\omega < \frac{\omega_{c}}{2}$}} {\sf (a) $E<V_{max}$} Here, depending on the signs of the various coefficients we will have two cases : (i) $\frac{mr^{2}\omega\omega_{c}}{2}\left (1+\frac{2\omega}{\omega_{c}} \right ) < E < V_{max}$ The solution here is : \begin{equation} \tan \left (\frac{\theta}{2} +\frac{\pi}{4} \right ) = \frac{k}{a} \frac{sn (X,k)}{dn(X,k)} \end{equation} where $\kappa = \frac{a}{\sqrt{a^{2}+b^{2}}}$, $a^{2}b^{2}=\alpha-1$, $a^{2}-b^{2}=2-\alpha-\beta$ and $X= \frac{a}{\kappa}\sqrt{\frac{E}{2mr^{2}}} (t-t_{0})$. (ii) $0<E<\frac{mr^{2}\omega\omega_{c}}{2}\left (1+\frac{2\omega}{\omega_{c}} \right )$ The solution here is the same as before except that now $a^{2}<b^{2}$ and therefore one may obtain it from the previous solution by just interchanging $a$ and $b$. {\sf (b) $E > V_{max}$} \begin{equation} \tan \left ( \frac{\theta}{2} +\frac{\pi}{4} \right ) = \frac{sn( X,k)}{acn(X,k)} \end{equation} where $ X=a{\sqrt{E \over 2mr^{2}}} (t-t_0)$, $a^2 b^2 = 1-\alpha$, $a^2+b^2 =2-\alpha-\beta$ \, . {\underline{\sf $\omega > \frac{\omega_{c}}{2}$}} {\sf (a) $E<V_{max}$} The two cases depending on the signs of the various coefficients are : (i) $0<E<mr^{2}\omega\omega_{c}$ Here the solution is : \begin{equation} \tan \left (\frac{\theta}{2}+\frac{\pi}{4} \right ) = \frac{k'^2}{b} \frac{sn(X,k)}{dn(X,k)} \, , \end{equation} where $a^{2}b^{2} = \alpha - 1$, $b^{2}-a^{2}=\alpha +\beta -2$, $\kappa =\frac{a}{\sqrt{a^{2}+b^{2}}}$ and of course $k'^2+k^2 =1$. (ii) $mr^{2}\omega\omega_{c}<E<V_{max}$ The solution is \begin{equation} \tan \left (\frac{\theta}{2} + \frac{\pi}{4} \right ) = \frac{1}{a} sn [a\sqrt{\frac{E}{2mr^{2}}} (t-t_{0} )] \, . \end{equation} {\sf (b) $E>V_{max}$} The two cases are given below : (i) $V_{max} < E < mr^{2}\left (\omega^{2}+\frac{\omega\omega_{c}}{2}\right )$ The solution is \begin{equation} \tan \left (\frac{\theta}{2}+\frac{\pi}{4} \right ) = \frac{1}{a} sn(X,k) \, . \end{equation} (ii) $E>mr^{2}\left (\omega^{2} + \frac{\omega\omega_{c}}{2} \right )$ The solution is \begin{equation} \tan \left (\frac{\theta}{2} + \frac{\pi}{4} \right ) = \frac{1}{a} \frac{sn(X,k)}{cn(X,k)} \, , \end{equation} where in both the above solutions $X=a\sqrt{\frac{E}{2mr^{2}}} (t-t_{0})$. Before concluding this section, we briefly discuss a couple of special cases in which the solution can be written in terms of the well known trigonometric and hyperbolic functions. {\underline{\sf Special solutions in the gravitational case} For $\omega < \omega_{0}$ if we choose $E= 2mr^{2}\omega_0^{2}$ then we obtain the following special solution \begin{equation} \tan \frac{\theta}{2} = \pm\frac{1}{a} cosech \left (\sqrt{\omega_{0}^{2} + \omega^{2}} (t-t_0)\right ) \, , \end{equation} where $a^2 = \frac{\omega_{0}^{2}}{\omega^{2}+\omega_{0}^{2}}$. This is known as the `sticking' solution--i.e. as $t\rightarrow \pm \infty,$ $\theta \rightarrow 0$ (maxima), while for $t\rightarrow t_{0},$ $\theta \rightarrow \pi$ (minima). If $\omega> \omega_{0}$, the same solution holds (with the same form of $a$) but at $E = V_{max} = {mr^2 \over 2\omega^2} (\omega^2 +\omega_0^2)^2$. It is also a sticking solution but it goes from local to absolute maxima as $t$ goes from $t_0$ to $\pm \infty$. It is the {\em value} of $\omega$ which makes the two solutions functionally different. {\underline{\sf Special solutions in the magnetic case} Here, for $\omega<\frac{\omega_{c}}{2}$ and $E = V_{max} =mr^{2}\omega\omega_{c}$ we find the following solution : \begin{equation} \tan ({\frac{\theta}{2} +\frac{\pi}{4}}) = \sqrt{\frac{1}{1-\frac{2\omega} {\omega_{c}}}} \sinh \left (\sqrt{\frac{\omega\omega_{c}}{2} \left (1-\frac{2\omega}{\omega_{c}} \right ) }(t-t_0)\right ) \end{equation} This is once again the so--called `sticking' solution--the particle reaches the maxima ($\pi /2$) as $t\rightarrow \pm \infty$ while it is at the minimum ($3\pi /2$) as $t\rightarrow t_0$. Another solution for $\omega<\frac{\omega_{c}}{2}$ is obtained at $ E=\frac{mr^{2}}{8} \left (\omega_{c} + 2\omega \right )^{2} > V_{max}$. This is given by \begin{equation} \tan \left ( \frac{\theta}{2} +\frac{\pi}{4} \right ) = \sqrt{\frac{ \omega_{c} +2\omega}{\omega_{c} - 2\omega}} \tan \left (\sqrt{1-\frac{4\omega^{2}}{\omega_{c}^{2}}} \frac{\omega_{c}}{4} (t-t_0) \right ) \, . \end{equation} This solution oscillates around the minima at $\theta =\frac {3\pi}{2}$. For $\omega>\frac{\omega_{c}}{2}$ and $E=mr^{2}\omega\omega_{c} < V_{max}$ we have a solution which oscillates around the minimum at $\theta=\frac{3\pi}{2}$. This is given by \begin{equation} y=\tan \left (\frac{\theta}{2} + \frac{\pi}{4} \right ) = \sqrt{\frac{\omega_{c}}{2\omega-\omega_{c}}} \sin \left (\sqrt{1- \frac{\omega_{c}}{2\omega}} \omega (t-t_{0}) \right ) \, . \end{equation} Finally, for $\omega > {\omega_c\over 2}$ and $E = {mr^2 \over 8} (\omega_c +2\omega)^2 = V_{max}$ we have the solution \begin{equation} \tan \left ( \frac{\theta}{2} +\frac{\pi}{4} \right ) = \sqrt{\frac{ 2\omega +\omega_{c}}{2\omega - \omega_{c}}} \tanh \left (\sqrt{1-\frac{\omega_{c}^{2}}{4\omega^{2}}} \omega (t-t_0) \right ) \, . \end{equation} This is again a sticking solution which goes from the absolute minimum to the maximum as $t$ goes from $t_0$ to $\pm \infty$. \section{Quantum mechanics of a PORC in a constant gravitational field--instantons} In the gravitational case, the potential has a pair of degenerate minima. Therefore, one has the possibility of constructing instanton solutions. Since the two minima are separated by two different barriers there are two types of instantons. In Appendix we also look at the possibility of constructing exact solutions of the Schr\"odinger equation in this case. \subsection{Instanton solutions} As mentioned earlier, the effective potential has degenerate minima and therefore we will obviously have instanton solutions. Since the minima are separated by two kinds of barriers we will have two different instantons --one across the barrier at $0 (2\pi)$ and another across the barrier at $\pi$. It is worth reminding that instantons are the finite Euclidean action solutions of the Euclidean equations of motion of any theory. Since we are dealing with particle mechanics we look at solutions to the Euclideanised Newton's second law. These solutions have the interpretation in terms of quantum tunneling. For a particle in a double well potential, classically, it is not possible to cross the barrier connecting the two vacua. However, when we Euclideanise the Newton's second law, we are essentially looking at the motion in Euclidean time in an inverted potential. These Euclidean solutions (called instantons) begin at one vacuum at $\tau\rightarrow -\infty$ and end at another at $\tau\rightarrow +\infty$. In the saddle--point approximation, the tunneling amplitude goes as $\exp(\frac{-S_{E}}{\hbar})$. The instanton solutions and aspects of quantum tunneling in this model have been worked out in an earlier paper by one of us \cite{sk:pla92}. We therefore only summarize these results here and refer the reader to that article for the relevant details. This section is included here entirely for the sake of completeness. Below we write down the instanton solutions and the corresponding Euclidean actions. To arrive at the instantons we need to solve the Euclidean equation of motion which is given by \begin{equation} \theta '' = - \omega^{2} (\cos \theta + a) \sin \theta \quad ; \quad a = {\omega_0^2 \over \omega^2} \, , \end{equation} where the prime denotes differentiation w.r.t. $\tau = -it$ (Euclidean time). It is easy to convert the above equation into one for $\theta '$ alone and then integrate the resulting first order equation. We first deal with the case of $\omega >\omega_{0}$. The instanton/anti--instanton and it's Euclidean action across the barrier at $\theta=2\pi(0)$ are given by \begin{equation}\label{t1} \theta_{1}(\tau) = \pm 2\tan^{-1}\left ( \sqrt{\frac{1+a}{1-a}}\tanh\left [ \sqrt{\frac{1}{4}(1-a^{2})}\omega\tau\right ]\right ) \, . \end{equation} \begin{equation} S_{E1}=4mg^{1/2}r^{3/2}\left [ \sqrt{\frac{1-a^{2}}{a}}+2\sqrt{a} \tan^{-1} \left (\sqrt{\frac{1+a}{1-a}} \right ) \right ] \, . \end{equation} The instanton/anti--instanton and it's Euclidean action across the barrier at $\theta=\pi$ are given by : \begin{equation}\label{t2} \theta_{2}(\tau) = \pm 2\tan^{-1}\left ( \sqrt{\frac{1+a}{1-a}}\coth\left [ \sqrt{\frac{1}{4}(1-a^{2})}\omega\tau\right ]\right ) \, . \end{equation} \begin{equation} S_{E2}=4mg^{1/2}r^{3/2}\left [ \sqrt{\frac{1-a^{2}}{a}}-2\sqrt{a} \tan^{-1} \left (\sqrt{\frac{1+a}{1-a}} \right ) \right ] \, . \end{equation} Notice that for $a\sim 1$ (i.e. $\omega \sim \omega^{0}$) the second term in each Euclidean action has the larger value. This implies that the Euclidean action is proportional to $\pm \frac{4\pi mgr}{\hbar \omega}$ and the tunneling amplitude will be dominated by the instanton across the $\theta =\pi$ barrier. On the other hand for $a\sim 0$ (i.e. $\omega \sim \infty $ or $r \rightarrow \infty $--very high frequencies or very large radius ) the Euclidean action is proportional to $\frac{4m\omega^{2}r}{\bar h \omega}$ and the contribution to the tunneling probablity from both the instantons are the same. In the latter case, the effect of gravity is washed out while in the former gravity predominates. For the $\omega <\omega_{0}$ case we have a single well in $S^{1}$ but infinite number of degenerate minima on the full line and the instanton/ anti--instanton as well as it's Euclidean action are given by \begin{equation}\label{t3} \theta_{3} (\tau) = \pm 2\tan^{-1}\left ( p \sinh (\omega_0 \tau \over p) \right ) \, . \end{equation} \begin{equation} S_{E3} =4mg^{1/2}r^{3/2}\left [ \sqrt{a}\tan^{-1}\left (\frac{1}{\sqrt{ a-1}} \right ) + \sqrt{\frac{a-1}{a}} \right ] \end{equation} with $p =\sqrt{\frac{a}{a-1}}$. For domains of $a\sim 1$ (when the first term in $S_{E3}$ dominates ) and $a >>1$ (when the second term in $S_{E3}$ gives the largest contribution) one can see the effects similar to the ones stated earlier. \section{Quantum mechanics of a PORC in a constant magnetic field --bounces, instantons and exact solutions} \subsection{ Exact Solutions} We now write down the Schr\"odinger equation for this system and look for exact solutions \cite{hks} following the method introduced by Razavy for bistable potentials {\cite{raz:ajp}}. For this case, the Schr\"odinger equation turns out to be \begin{equation}\label{se} \frac{d^{2}\Psi}{dx^{2}} + \left [ \frac{8mr^{2}E}{{\hbar}^{2}} + \frac{2m^{2}r^{4}\omega^{2}}{{\hbar}^{2}} - \frac{4m^{2}r^{4}\omega \omega_{c}}{{\hbar}^{2}} \cos 2x + \frac{2m^{2}r^{4} \omega^{2}} {{\hbar}^{2}} \cos 4x \right ] \Psi = 0 \, , \end{equation} where we have introduced the following redefinitions $\theta --> \frac{\pi}{2} - \theta$ and then $\theta --> 2 x$. We further rewrite the Schr\"odinger equation in the following form : \begin{equation} \frac{d^{2} \Psi}{dx^{2}} + \left [ \bar \epsilon -(n+1)\xi \cos 2x + \frac{1}{8} \xi^{2} \cos 4 x \right ] \Psi = 0 \end{equation} where $ \epsilon = \frac{8mr^{2}E}{\hbar^{2}} \quad ; \quad \bar \epsilon = \epsilon + \frac{1}{8}\xi^{2} \quad ; \quad (n+1) = \frac{mr^{2}\omega_{c}}{\hbar} \quad ; \quad \xi = {4m r^2 \omega \over \hbar}$ . We shall now write down the exact solutions for different values of $n$ which, of course, correspond to different Lagrangians/Hamiltonians with the same functional form. In fact, as is clear from the definition given above, $n$ is related to the ratio of the {\em critical} angular momentum $mr^{2}\omega_{c}$ and Planck's constant $\hbar$. Following Razavy, we first write down a solution for $n=0$ (and also $\bar \epsilon =0$) and then define the wave functions for $n>0$ as the product of the $n=0$ solution and an unknown function $\Phi (x)$. We thus start with the ansatz \begin{equation} \Psi (x) = \exp \left ( -\frac{\xi}{4}\cos 2x \right ) \Phi (x) \, . \end{equation} The resulting differential equation for $\Phi (x)$ turns out to be \begin{equation} \Phi^{''} + \xi \sin 2x \Phi^{'} + \left ( \bar \epsilon + {\xi^2 \over 8} - n \xi \cos 2x \right ) \Phi = 0 \end{equation} with the primes denoting differentiation w.r.t. $x$. It is easily shown that this is a quasi-exactly solvable (QES) system \cite{km}. In particular, one can easily show that the above equation is equivalent to a three term difference equation which is exactly solvable in case $n$ is a non-negative integer. In particular, for $n= 0,1,2,...$, first $(n+1)$ solutions of period $\pi$ ($2\pi$) are obtained in case $n$ is even (odd). For example, the solutions for $n= 0,1,2,3$ are ${\bf n = 0 :}$ \centerline{ $ \Phi_{0} = 1 \quad ; \quad $ $E_{0} = -{1\over2}\hbar \omega \frac{\omega}{\omega_{c}}$} \, . ${\bf n = 1 :}$ \centerline{$\Phi = \sin x \quad ; \quad $ $ E = \frac{1}{16}\hbar \omega_{c} - \frac{1}{2} \hbar \omega \left ( \frac{1 + 2\omega}{\omega_{c}} \right ) $ } \, . \centerline{$\Phi = \cos x \quad ; \quad $ $ E = \frac{1}{16}\hbar \omega_{c} + \frac{1}{2} \hbar \omega \left (1-2\frac{\omega}{\omega_{c}} \right ) $} \, . ${\bf n = 2 :} $ \centerline{ $\Phi_{0} = 2\xi +a_{-} \cos 2x \quad ; \quad $ $ E_{0} = \frac{\hbar \omega_{c}}{12} -\frac{3}{2} \hbar \omega \frac{\omega}{\omega_{c}} - \frac{\hbar\omega_{c}}{12} \sqrt{1 + 144\frac{\omega^{2}}{\omega_{c}^{2}}}$ } \, . \centerline{$\Phi = \sin 2x \quad ; \quad $ $E = \frac{1}{6}\hbar \omega_{c} - \frac{3}{2} \hbar \omega \frac{\omega}{\omega_{c}}$ } \, . \centerline{$\Phi = 2\xi +a_{+}\cos 2x \quad ; \quad $ $ E = \frac{\hbar \omega_{c}}{12} -\frac{3}{2} \hbar \omega \frac{\omega}{\omega_{c}} + \frac{\hbar\omega_{c}}{12} \sqrt{1 + 144\frac{\omega^{2}}{\omega_{c}^{2}}}$} \, . $\bf n = 3: $ \centerline{$\Phi = 3\xi \cos x + b_{0}\cos 3x \quad ; \quad $ $E = \frac{\hbar \omega_{c}}{32} \left ( -\frac{1}{4}\xi^{2} + 5 + \xi -2 \sqrt{\xi^{2} +4 - 2\xi} \right )$} \, . \centerline{$\Phi = 3\xi \sin x + b_{1}\sin 3x \quad ; \quad $ $E = \frac{\hbar \omega_{c}}{32} \left ( -\frac{1}{4}\xi^{2} + 5 - \xi -2 \sqrt{\xi^{2} +4 + 2\xi} \right )$} \, . \centerline{$\Phi = 3\xi \cos x + b_{2}\cos 3x \quad ; \quad $ $E = \frac{\hbar \omega_{c}}{32} \left ( -\frac{1}{4}\xi^{2} + 5 + \xi +2 \sqrt{\xi^{2} +4 - 2\xi} \right )$} \, . \centerline{$\Phi = 3\xi \sin x + b_{3}\sin 3x \quad ; \quad $ $E = \frac{\hbar \omega_{c}}{32} \left ( -\frac{1}{4}\xi^{2} + 5 - \xi +2 \sqrt{\xi^{2} +4 + 2\xi} \right )$} \, . where $a_{\pm} = 2 \pm 2 \sqrt{1+\xi^2}$ \, ; \ $ b_{0,2} = 4-\xi \pm 2\sqrt{\xi^{2}-2\xi + 4} \, ; \ b_{1,3} = 4+\xi \pm 2\sqrt{\xi^{2} + 2\xi + 4}$ \, . A remark is in order at this stage. Since we are treating the problem of a point particle on circle $S^{1}$, and since the potential in the Schr\"odinger eq. (\ref{se}) satisfies the boundary condition \begin{equation} V(x+\pi) = V(x) \, , \end{equation} hence physical considerations demand that the corresponding wave functions must also satisfy the boundary condition \begin{equation} \psi(x+\pi) = \psi(x) \, . \end{equation} In that case, the solutions obtained for $n=1,3$ are unacceptable as they do not satisfy this boundary condition but rather they change sign under $x \rightarrow x+\pi$. However, if we are considering it as a periodic problem on the full line then of course these are acceptable solutions. We may add that the ground state energies obtained here (for $n = 0,2$) are useful in another context. In particular, it has been shown using the transfer integral method \cite{ssf} that the classical free energy of the soliton bearing field theories at low temperatures is given by the ground state energy of the corresponding Schr\"odinger like equation. Thus the ground state energies obtained here for $n=0,2$ are of direct relevance in the context of the classical free energy of the corresponding double sine-Gordon field theory. \subsection{ Euclidean time solutions--Bounces and Instantons} Since the effective potential also has a false vacuum, we expect that `bounce' solutions exist to the classical equations of motion in Euclidean time. The classical equation of motion in Euclidean time which we solve is given by \begin{equation} {\theta '}^{2} = \omega^{2} \left [ (1-\sin \theta) (1+\sin\theta - a) \right ] \, , \end{equation} where we have added a constant to the original effective potential in order to write it in the above form (note that this does not effect the classical equations of motion which we intend to solve). Here $a = {\omega_c \over \omega} < 2$. However, this choice of the effective potential forces us to confine ourselves to $a<2$. For $a=2$ the equation is meaningless. A straightforward integration of the above equation yields the following solutions : \begin{equation}\label{t4} \theta_{1} (\tau) = \frac{\pi}{2} + 2\tan^{-1}\left [ \frac{1}{\alpha} sech \sqrt{\omega(\omega-\frac{\omega_{c}}{2})} \tau \right ] \, . \end{equation} \begin{equation}\label{t5} \theta_{2} (\tau) = \frac{5\pi}{2} - 2 \tan^{-1}\left [ \frac{1}{\alpha} sech \sqrt{\omega (\omega-\frac{\omega_{c}}{2})} \tau \right ] \, . \end{equation} The first of these is the bounce across the barrier at $\sin^{-1}\frac{\omega _{c}}{2\omega}$ while the second one is across the barrier at $\pi - \sin^{-1} \frac{\omega_{c}}{2\omega}$. Note that the Euclidean actions for both these solutions are the same--this is because the barrier heights which separate the false and the true vacua are the same. The expression for the Euclidean action turns out to be \begin{equation} S_{E} = 2mr^{2}\omega_{c}\sqrt{1-\frac{a}{2}} \left [ \frac{2+a}{a} - \sqrt{\frac{2-a}{2}} \sinh^{-1}\sqrt{\frac{2-a}{2}} \right ] \end{equation} where $\alpha ^{2} = \frac{a}{2-a}$. Using the Euclidean action one can now evaluate the decay rate of the false vacuum by the formula $\Gamma \sim \exp \left (-\frac{S_{E}}{\not h}\right )$. For a periodic potential generalization of the magnetic case problem one has degenerate absolute minima at ${(2n+1)\pi \over 2}$ and one can obtain an `instanton' which starts out at, say, $\theta =-\frac{\pi}{2}$ (as $\tau \rightarrow -\infty$) and crosses the local minimum at $\theta = \frac{\pi}{2}$ (as $\tau \rightarrow 0$) to end up at $\theta=\frac {3\pi}{2}$ (as $\tau \rightarrow \infty$). To that end, let us first note that by adding a suitable constant one can also write the classical equation of motion in Euclidean time in an alternative form as \begin{equation} {\theta '}^{2} = \omega^{2} \left [ (1+\sin \theta) (1-\sin\theta + a) \right ] \, . \end{equation} It is now easily shown that irrespective of whether $\omega >$ or $<\omega_c/2$, the instanton solution is given by \begin{equation}\label{t6} \theta_3 = \frac{\pi}{2} + 2\tan^{-1} \left [ \frac{1}{\sqrt{1+\frac{2\omega}{ \omega_{c}}}}\sinh \left ( \sqrt{(\omega_{c}+2\omega)2\omega} \frac{\tau}{2} \right ) \right ] \, , \end{equation} with the corresponding action being \begin{equation} S_E = -{1\over 2a^2} + {1\over 4a^3} {(4a^2-1)\over \sqrt{a^2 -1}} \ln \bigg [ {a+\sqrt{a^2 -1} \over a-\sqrt{a^2-1}} \bigg ] \, . \end{equation} where $a = \sqrt{1+{2\omega \over \omega_c}}$. \section{Small oscillations about instantons and bounces-- qualitative analysis} In this section we present a qualitative analysis of the problem of small oscillations about an instanton/bounce solution. We do {\em not} explicitly solve the corresponding Schr\"odinger-like equations but discuss qualitatively the nature of the effective potentials, the possible existence of negative eigenvalues and bound states. It may however be added that in all the cases we have been able to reduce the stability equation to Heun's equation. Additionally, the analysis for the small oscillations about the DSG kink has been discussed in \cite{osc:ref}. The equation governing small oscillations is given by: \begin{equation} \left (-\partial_{\tau}^{2} + V''(\theta (\tau))\right ) \chi_{n}= \lambda_{n} \chi_{n} \, . \end{equation} This is a `time--independent' Schr\"odinger equation with the potential $V''(\theta (\tau))$. We therefore need to solve the corresponding eigenvalue problem and look for the existence of a negative eigenvalue which will indicate an instability. Below we write down the `potential' in the two cases of the gravitational and magnetic fields. {\bf 1. : System in a gravitational field} Since for $\omega>\omega_{0}$ we have two types of instantons we need to evaluate $V ''(\theta)$ for each of these cases separately. (a) Using eq. (ref{t1}), $V''(\theta)$ for the instanton across the $2\pi$ barrier turns out to be \begin{equation} V''(\theta_1 (\tau)) = -mr^{2}{\omega}^{2}(1-a^2)\frac{\left [(1-a) + (1+a) \tanh^{4}{b\omega\tau}-6\tanh^{2}b\omega\tau \right ]}{\left [ (1-a)+(1+a)\tanh^{2}b\omega\tau \right ]^{2}} \, . \end{equation} This is shown in Fig 5. (b) Using eq. (\ref{t2}), $V''(\theta (\tau ))$ for the instanton across the $\pi$ barrier turns out to be \begin{equation} V''(\theta_2 (\tau)) = -mr^{2}{\omega}^{2}(1-a^2)\frac{\left [(1-a) + (1+a) \coth^{4}{b\omega\tau}-6\coth^{2}b\omega\tau \right ]}{\left [ (1-a)+(1+a)\coth^{2}b\omega\tau \right ]^{2}} \, . \end{equation} This is plotted in Fig 6. A condition for the stability of a given instanton is the existence of a node-less zero--mode solution (solution with a zero eigenvalue) to the Schr\"odinger equation governing small oscillations. The zero mode solution can be written down very easily-- $\chi_{0}$ is just equal to $\frac{d\theta(\tau)}{d\tau}$ (details are there is [5], [6] and [7]). For the instanton across $\theta = 2\pi$, using eq. (\ref{t1}), one finds that \begin{equation} \chi_0 = \frac{d\theta_{1}(\tau)}{d\tau} = (1-a^{2}) \omega \frac{1} {[2\cosh^{2}\beta \tau -(1+a)`]} \end{equation} where $\beta = \sqrt{\frac{1-a^{2}}{2}} \omega$. Note that this solution is node-less which implies that the corresponding instanton is stable. On the other hand, it can be readily shown that the instanton across the $\theta=\pi$ barrier is also stable. In particular, using eq. (\ref{t2}), the zero--mode solution is given by \begin{equation} \chi_{0} = \frac{d\theta_{2}(\tau)}{d\tau} = -(1-a^{2}) \omega \frac{1}{[(1+a) +2\sinh^{2}\beta \tau]} \, . \end{equation} which is clearly nodeless as well. In the case of $\omega <\omega_0$ there is only one instanton and using eq. (\ref{t3}), $V''(\theta)$ turns out to be \begin{equation} V''(\theta_3 (\tau)) = mr^2 \omega^2 \frac{\bigg [(a-1)p^4 \sinh^4 y +6p^2 \sinh^2 y - (a+1) \bigg ]}{(1+p^2 \sinh^2y)} \, , \end{equation} and the corresponding zero-mode solution is given by \begin{equation} \chi_0 = \frac{d\theta_{3}(\tau)}{d\tau} = \frac{2\omega_0 \cosh y} {1+p^2 \sinh^2 y} \, , \end{equation} which is clearly node-less. Here $y = {\omega_0 \tau \over p}$ with $a= {\omega_0^2 \over \omega^2}$ and $p = \sqrt{{a \over a-1}}$. {\bf Case 2: System in a magnetic field} For the bounce solution in the PORC in a magnetic field, using eq. (\ref{t4}) (or (\ref{t5})), the potential $V''(\theta)$ in the Schr\"odinger-like equation turns out to be \begin{equation} V''(\theta_{1,2}(\tau)) = \left (\frac{mr^2{\omega}^2}{2-a} \right ) \frac{\left [-a^2\cosh^{4}\beta \tau + 8a\cosh^{2}\beta \tau - (4-a^2) \right ]} {(1+ \alpha^2 \cosh^{2}\beta \tau)^2} \, . \end{equation} This is plotted in Fig.7. The corresponding zero-mode solution is given by \begin{equation} \chi_0 = \frac{d\theta_{1,2}(\tau)}{d\tau} = \alpha \omega \sqrt{2(2-a)} \left (\frac{\sinh{y}}{1+ \alpha^2 \cosh^2{y} } \right ) \, , \end{equation} which clearly has a node at $\tau=0$. Here $y = {(2-a)\omega \tau \over 2}$. On the other hand, corresponding to the instanton solution in a magnetic field, using eq. (\ref{t6}), $V''(\theta)$ turns out to be \begin{equation} V''(\theta_{3}(\tau) = {m r^2 \omega^2 (a+1) \over a} {\bigg [a(6+a) -2(1+3a) \cosh^2{y} + \cosh^4{y} \bigg ] \over (a + \cosh^2{y})^2} \, . \end{equation} The corresponding zero-mode solution is given by \begin{equation} \chi_0 = \frac{d\theta_{3}(\tau)}{d\tau} \propto \left ({\cosh{y} \over a + \cosh^2{y}} \right ) \, , \end{equation} which clearly is node-less. Here $a = {2\omega \over \omega_c}, \ y = \sqrt{a(a+1)} {\omega_c \tau \over 2}$. \section{Summary and outlook} We have discussed a variety of aspects of the toy model of a particle on a rotating loop. After defining a generic class of models (via a choice of a class of trigonometric potentials ) we identified two specific cases. These included the effects of a (i) uniform gravitational field and (ii) a uniform magnetic field. The classical mechanics for both these cases was worked out. Subsequently we dealt with the quantum mechanical problem in either case from the (i) exact as well as the (ii) semi-classical standpoints. For the gravitational case we found exact solutions by identifying the time--independent Schr\"odinger equation with the Whittaker--Hill equation. These solutions were {\em formal} in the sense that the energy eigenvalues remain hidden in the continued fractions. We have not attempted to arrive at the expressions for the energy eigenvalues by solving the continued fractions. Hence this discussion appears in the Appendix to the paper. Apart from the exact solutions we have also looked at the instantons in this model. We constructed the instantons and the corresponding Euclidean actions which gave us a feeling for the quantum tunneling phenomena in this example. In the magnetic case the potential also has a false vacuum. Further, for special values of the parameters we were able to obtain some of the energy eigenvalues and eigenfunctions. Thereafter, we constructed the corresponding bounce (instanton) solutions and their Euclidean actions to arrive at the decay rate of the false vacuum (feel for quantum tunneling phenomena). We then looked into the problem of small oscillations about the instantons and bounces by investigating the nature of the potential in the Schr\"odinger--type equation which govern these perturbations. This section is largely qualitative. Even though we did not solve the relevant Schr\"odinger-like equation, we have shown that in all the cases the Schr\"odinger equation reduces to Heuns equation which is known to have four regular singular points. Further, as expected we have seen that the zero mode is stable (unstable) depending on whether one is considering an instanton (bounce) solution. Finally, the connection of these toy examples with several models in field theory and statistical mechanics was pointed out. The aim of the paper has been to point out the diverse aspects of the problem of a particle on a rotating loop. It is quite illuminating that a lot of important concepts in field theory can be illustrated through the study of this simplistic example. We conclude by pointing out some other aspects which may be dealt with in future. \begin{itemize} \item{ Aspects of chaos in PORC with a $\delta$ function kick--the generalization of the kicked pendulum/kicked rotor} \item{Investigating what happens at finite temperature by utilizing techniques of finite temperature quantum mechanics/quantum field theory} \item{Exact classical and quantum solutions for the model with a gravitational, an electric and a magnetic field all put in together} \end{itemize} \section*{Acknowledgements} SK thanks IUCAA, Pune for financial support during the period when this work began. \newpage \centerline{{\bf APPENDIX : Exact solutions for PORC in a gravitational field}} \vspace{.2in} The time-independent Schr\"{o}dinger equation for this system is given by \begin{equation} \frac{d^{2}\psi}{d\theta^{2}}+\frac{2m}{\hbar^{2}}\left[ E+\frac{1}{4}mr^{2}\omega^{2}-mgr\cos\theta-\frac{1}{4}mr^{2}\omega^{2} \cos2\theta\right]\psi=0 \, . \end{equation} Introducing a new variable $\alpha=\theta/2$ we arrive at the following equation \begin{equation}\label{c3} \frac{d^{2}\psi}{d\alpha^{2}}+\left[\left(\frac{8mE}{\hbar^{2}}+ \frac{2m^{2}r^{2}\omega^{2}}{\hbar^{2}}\right)- \frac{8m^{2}gr}{\hbar^{2}}\cos2\alpha-\frac{2m^{2}r^{2}\omega^{2}} {\hbar^{2}}\cos4\alpha\right]\psi=0 \, . \end{equation} This equation is the Whittaker-Hill equation.We now digress briefly to a discussion of the solutions of this equation. In the theory of periodic differential equations, the Whittaker-Hill equation is a special case of the general Hill equation given by \begin{equation} \frac{d^{2}\psi}{d\theta^{2}}+\left(\sum_{n=0}^{\infty}A_{n}\cos 2n\theta\right)\psi=0 \, . \end{equation} Whittaker-Hill equation is a special case of this general Hill equation and is also known as the three term Hill equation, i.e. we have only the $n=0,1,2$ terms remaining in the infinite series given above. Thus we have \begin{equation}\label{8.1} \frac{d^{2}\psi}{d\theta^{2}}+\left(A_{0}+A_{1}\cos2\theta+ A_{2}\cos4\theta\right)\psi=0 \, . \end{equation} Such equations arise when the Helmholtz equation $\left(\nabla^{2}+k^{2}\right) \psi=0$ is separated in a general paraboloidal coordinate system. We shall be concerned here with the form of equation for which $A_{1}$ and $A_{2}$ are both negative.Following \cite{whe:ref} we define \begin{equation} |A_{2}|=\frac{1}{8}\omega^{2},A_{0}=\eta+\frac{1}{8}\beta^{2}, A_{1}=-\rho\beta \, . \end{equation} Hence we have \begin{equation}\label{c4} \frac{d^{2}\psi}{d\theta^{2}}+\left(\eta+\frac{1}{8}\beta^{2}- \rho\beta\cos2\theta-\frac{1}{8}\beta^{2}\cos4\theta\right)\psi=0 \, . \end{equation} The transformation $\psi=\phi\exp(-\frac{1}{4}i\beta\cos2\theta)$ reduces (\ref{8.1}) to Ince's equation \begin{equation} \frac{d^{2}\phi}{d\theta^{2}}+i\beta\sin2\theta\frac{d\phi}{d\theta} +\left[\eta+\beta(\rho+i)\cos2\theta\right]\phi=0 \, . \end{equation} Solutions to the above equation falls into four classes depending on whether the function $\phi$ is even or odd and has period $\pi$ or $2\pi$. These solutions are in terms of infinite Fourier series with the relation between the coefficients of the series given in terms of three term recurrence relations. The recurrence relations can be reduced to an infinite continued fraction which is known as the corresponding characteristic equation for the relevant function. We shall primarily be concerned with solutions of period $\pi$. This is because the equation governing our physical problem is written in terms of a new variable $\alpha= \theta/2$. Thus $\pi$ periodic solution in $\alpha$ will turn out to be $2\pi$ periodic solutions in $\theta$. We now list the even and odd $\pi$ periodic solutions of the Whittaker-Hill equation and the relevant continued fractions. Details about how these are obtained and a more thorough treatment of various issues related to periodic differential equations can be found in \cite{whe:ref}. \subsection{Even Solutions of Period $\pi$} \begin{equation} \phi(\theta)=\sum_{r=0}^{\infty}A_{2r}\cos2r\theta\equiv gc_{2n}(\theta, \omega,\rho) \, ; \ A_{2r} = X_{2r}B_{2r} \, , \end{equation} where \begin{equation} X_{2r}=(-\rho+i)(-\rho+3i)\cdots\left[-\rho+(2r-1)i\right] \, ; r\geq1 \, , \ X_{0}=1 \, , \end{equation} and \begin{equation} \frac{B_{2r}}{B_{2r+2}}=\frac{\frac{1}{2}\beta}{4r^{2}-\eta-} \frac{\frac{1}{4}\beta^{2}\{\rho^{2}+(2r+1)^{2}\}}{4(r+1)^{2}-\eta-} \frac{\frac{1}{4}\beta^{2}\{\rho^{2}+(2r+3)^{2}\}}{4(r+2)^{2}-\eta-} \cdots \, . \end{equation} The Continued fraction for characteristic values is given by \begin{equation}\label{c1} -\frac{2\eta}{\beta(\rho^{2}+1)}=\frac{\beta}{4-\eta-} \frac{\frac{1}{4}\beta^{2}(\rho^{2}+9)}{16-\eta-} \frac{\frac{1}{4}\beta^{2}(\rho^{2}+25)}{36-\eta-}\cdots \, . \end{equation} \subsection{Odd Solutions of Period $\pi$} \begin{equation} \phi(\theta)=\sum_{r=1}^{\infty}C_{2r}\sin2r\theta\equiv gs_{2n+2}(\theta, \omega,\rho) \, ; \ C_{2r} = Y_{2r} D_{2r} \, , \end{equation} where \begin{equation} Y_{2r}=(-\rho+3i)(-\rho+5i)\cdots\left[-\rho+(2r-1)i\right] \, , \ r\geq2 \, , \ Y_{2}=1 \, , \end{equation} and \begin{equation} \frac{D_{2r}}{D_{2r-2}}=\frac{\frac{1}{2}\beta}{4r^{2}-\eta-} \frac{\frac{1}{2}\beta^{2}\{\rho^{2}+(2r+1)^{2}\}}{4(r+1)^{2}-\eta-} \frac{\frac{1}{2}\beta^{2}\{\rho^{2}+(2r+3)^{2}\}}{4(r+2)^{2}-\eta-} \cdots \, . \end{equation} The Continued fraction for characteristic values is given by \begin{equation}\label{c2} \frac{4-\eta}{\frac{1}{2}\beta(\rho^{2}+9)}=\frac{\frac{1}{2}\beta}{16-\eta-} \frac{\frac{1}{4}\beta^{2}(\rho^{2}+25)}{36-\eta-} \frac{\frac{1}{4}\beta^{2}(\rho^{2}+49)}{64-\eta-}\cdots \, . \end{equation} Equations (\ref{c1}) and (\ref{c2}) are the equations which will give us the energy levels. Comparing eqs. (\ref{c3}) and (\ref{c4}) we find that \begin{equation} \eta=\frac{8mEr^{2}}{\hbar^{2}} \, , \ \frac{1}{8}\beta^{2} =\frac{2m^{2}r^{4}\omega ^{2}}{\hbar^{2}} \, , \ \rho=\frac{2mgr}{\hbar\omega} \, . \end{equation} Fixing $m,r$ and $\omega$ we can evaluate the energy levels of the system by solving the respective characteristic equations (for even and odd functions) iteratively.
\section{Introduction} \subsection{} Let $k$ be a field of characteristic zero, $\CO$ be a dg operad over $k$ and let $A$ be an $\CO$-algebra. In this note we define formal deformations of $A$, construct the deformation functor $$\Def_A:\dgar(k)\to\simpl$$ from the category of artinian local dg algebras (see~\ref{notation} for the precise definition) to the category of simplicial sets. In the case $\CO$ and $A$ are non-positively graded, we prove that $\Def_A$ is governed by the tangent Lie algebra $T_A$ defined in~\cite{haha}. A very easy example~\ref{void} shows that the result does not hold without this condition. \subsection{} ``Classical'' formal deformation theory over a field of characteristic zero can be described as follows. Let $k$ be a field of characteristic zero, $\art(k)$ be the category of artinian local $k$-algebras with residue field $k$. Let $\CC$ be a category cofibred over $\art(k)$. Equivalently, this means that a 2-functor $$ \CC:\art(k)\to \Cat$$ is given, that is a collection of categories $\CC(R),\ R\in\art(k)$, of functors $f^*:\CC(R)\to\CC(S)$ for each morphism $f:R\to S$ in $\art(k)$ and of isomorphisms $f^*g^*\iso (fg)^*$ satisfying the cocycle condition. Finally, let an object $A\in\CC(k)$ be given. Then the deformation functor $$\Def_A:\art(k)\to\Grp$$ assigns to each $R\in\art(k)$ the groupoid whose objects are isomorphisms $\alpha:\pi^*(B)\to A$ where $\pi:R\to k$ is the natural map, and morphisms are isomorphisms $B\to B'$ compatible with $\alpha$ and $\alpha'$. If we are lucky enough (and this is usually the case), one can naturally assign to $A$ a dg Lie algebra $\fg$ which governs deformation of $A$ in the following sense: there is a natural equivalence of groupoids $$ \Def_A\to\Del_{\fg}$$ as functors from $\art(k)$ to $\Grp$, where $\Del_{\fg}$ is the Deligne groupoid of $\fg$ -- see~\cite{gm,ddg}. As we argued in~\cite{uc}, the existence of dg Lie algebra $\fg$ is equivalent to the existence of formal dg moduli --- given by the dg coalgebra $\CC(\fg)$. This also means that the functor $\Def_A$ can be extended to a functor on the category $\dgar(k)$ of non-positively graded dg commutative artinian algebras with residue field $k$ with values in $\simpl$. \subsection{}The category of algebras over a given dg operad $\CO$ is not just a category --- there exist weak equivalences, homotopies and higher homotopies between the algebras. Therefore, the above described approach can not produce a reasonable definition of formal deformations of operad algebras. Also, we want that deformations of quasi-isomorphic algebras be equivalent, as well as deformations of algebras over quasi-isomorphic operads, so we should use a sort of ``derived'' notion of the deformation groupoid. \subsubsection{} Morally, the picture should be the following. For each $R\in\dgar(k)$ a $\infty$-category of $R\otimes\CO$-algebras should be defined; denote it $\Alg^{\infty}(\CO,R)$. The collection of $\Alg^{\infty}(\CO,R)$ should form an $\infty$-category cofibred over $\dgar(k)$. Let now $A\in\Alg^{\infty}(\CO,k)$. Then the deformation functor $$\Def_A:\dgar(k)\to\Grp^{\infty}$$ should be a ($\infty$-) functor to $\infty$-groupoids; its objects are $\infty$-isomorphisms $\alpha:\pi_*(B)\to A$ and morphisms --- $\infty$-isomorphisms $B\to B'$ commuting with $\alpha$ and $\alpha'$. \subsubsection{} Since we do not know well what an $\infty$-category is and how to assign an $\infty$-category to the category of operad algebras, we are looking for an appropriate substitute. According to ~\cite{haha}, the category $\Alg(\CO,R)$ of $R\otimes\CO$-algebras admits a simplicial closed model category (SCMC) structure. As a substitute to the $\infty$-category $\Alg^{\infty}(\CO,R)$, we suggest the simplicial category $\Alg^c_*(\CO,R)$ of cofibrant $R\otimes\CO$-algebras. This allows us to define a deformation groupoid as a functor \begin{equation} \Def_A:\dgar(k)\to\simpl \label{def-fr-eq} \end{equation} to the simplicial sets. Now, according to the general philosophy of deformation theory, the functor $\Def_A$ should be equivalent to the nerve of a certain dg Lie algebra $\fg$ --- see~\cite{uc}, Sect.~8 --- which is ``responsible'' for the deformations of $A$. In~\cite{haha}, Sect.~7, we constructed a functorial tangent dg Lie algebra $T_A\in\dgl(k)$ as the Lie algebra $\Der(\widetilde{A},\widetilde{A})$ of derivations of a cofibrant resolution $\widetilde{A}$ of $A$. The main result of this paper says that $T_A$ governs the formal deformations of $A$ provided $\CO$ and $A$ belong to $C^{\leq 0}(k)$. Quite unexpectedly, there is a very simple counter-example showing that the last condition is necessary --- see Example~\ref{void}. \subsection{Content of the Sections} Throughout the paper we work a lot with simplicial categories and simplicial groupoids. We collect in the Appendix the necessary information about the subject. It is mostly well-known or easily imaginable; thus, the closed model category structure on simplicial categories is a slight generalization of a result of~\cite{dk}. In Section~\ref{def-fr} we construct the deformation functor~ (\ref{def-fr-eq}). To compare it with the nerve of the tangent dg Lie algebra, we provide in Section~\ref{newnerve} a version of the nerve construction of ~\cite{uc}, Sect. 8, which assigns to a dg Lie algebra $\fg$ and to a dg artinian algebra $R$ a simplicial groupoid. Finally, in Section~\ref{final} we prove the main result. It follows easily from the model category structure of the category of simplicial categories. \subsection{Notation} \label{notation} In what follows we use the following notations for different categories. $\Ens,\Grp,\Cat$ are the categories of sets, small groupoids and small categories respectively. $\Delta$ is the category of ordered sets $[n]=\{0,\ldots,n\}, \ n\geq 0$ and order-preserving maps. For a category $\CC$ we denote by $\Delta^0\CC$ the category of simplicial objects in $\CC$. A simplicial category (and, in particular, a simplicial groupoid) will be supposed to have a discrete set of objects, if it is not explicitly specified otherwise. The category of small simplicial categories is defined $\sCat$ and that of simplicial groupoids $\sGrp$. For a fixed field $k$ of characteristic zero $\dgl(k)$ is the category of dg Lie algebras and $\dgar(k)$ is the category of non-positively graded commutative artinian dg algebras with residue field $k$. \subsection{Acknowledgement}This work was made during my stay at the Max-Planck Institut f\"ur Mathematik at Bonn. I express my gratitude to the Institute for the hospitality. \section{Homotopy algebras} \label{def-fr} In this Section we define deformations of dg operad algebras. Let $k$ be a fixed field of characteristic zero, $\CO\in\Op(C(k))$ be a dg operad over $k$, $A\in\Alg(\CO)$ be a $\CO$-algebra. Let $R$ be a commutative dg $k$-algebra. Componentwise tensoring by $R$ defines a functor on $\Op(C(k))$ (its values can be equally considered in $\Op(C(k))$ and in $\Op(\Mod(R))$). We define by $\Alg(\CO,R)$ the category of $R\otimes\CO$-algebras. This category admits a CMC structure --- see~\cite{haha} with quasi-isomorphisms as weak equivalences and surjective maps as fibrations. We denote by $\Alg^c(\CO,R)$ the full subcategory of cofibrant algebras and by $\CW^c(\CO,R)$ the category of cofibrant $\CO$-$R$-algebras with weak equivalences as arrows. The category $\Alg(\CO)$ admits also a simplicial structure so that Quillen's axiom (SM7) is satisfied --- see~\cite{haha}, 4.8. The simplicial structure is defined by the simplicial path functor which assigns to an algebra $A\in\Alg(\CO)$ and to a finite simplicial set $S\in\simpl$ the algebra $A^S=\Omega(S)\otimes A$ where $\Omega(S)$ denotes the dg commutative algebra of polynomial differential forms on $S$. In the sequel we will add asterisk to denote that we consider the corresponding simplicial category. Thus, for example, $\CW^c_*(\CO,R)$ is the simplicial category whose objects are cofibrant $R\otimes\CO$-algebras and whose $n$-morphisms from $x$ to $y$ consist of quasi-isomorphisms $x\to \Omega(\Delta^n)\otimes y$. Recall~\cite{uc} that it is worthwhile to consider artinian local non-positively graded dg $k$-algebras $(R,\fm)\in\dgar(k)$ as bases of formal deformations. \subsection{} \begin{defn}{dha}Let $A\in\Alg(\CO)$. {\em Deformation functor} $$\Def_A:\dgar(k)\to\simpl$$ is defined by the formula \begin{equation} \Def_A(R)=\hfib_A\left(\CN(\CW^c_*(\CO,R))\to\CN(\CW^c_*(\CO,k))\right). \label{def-def-functor} \end{equation} Here $\CN:\sCat\to\simpl$ is the simplicial nerve functor (see ~\ref{snerve}) and the homotopy fiber $\hfib$ being taken at a point $\widetilde{A}\in\CW^c_*(\CO,k)$ where $\widetilde{A}\to A$ is a cofibrant resolution of $A$. \end{defn} \subsubsection{}Recall (see~\cite{haha}) that for a $\CO$-algebra $A$ its tangent Lie algebra $T_A$ is defined as $$ T_A=\Der(\widetilde{A},\widetilde{A}).$$ Now we are ready to formulate the main result of this note. \subsubsection{} \begin{thm}{main} Let $\CO$ be a dg operad over a field $k$ of characteristic zero and let $A$ be an $\CO$-algebra. Suppose that both $\CO$ and $A$ are non-positively graded. Then the deformation functor $\Def_A:\dgar(k)\to\simpl$ is equivalent to the nerve $\Sigma_{\fg}$ of the tangent dg Lie algebra $\fg:=T_A$. \end{thm} \Thm{main} will be proven in Section~\ref{final}. Now we give an elementary example which shows that the non-positivity condition is necessary. \subsection{Example} \label{void} Let $\CO$ be the trivial operad $\CO(1)=k\cdot 1,\ \CO(i)=0$ for $i\not=0$. $\CO$-algebras are just complexes and derivations are just all endomorphisms. Let $A$ be the complex with zero differential with $A^i=k$ for all $i\in\Bbb{Z}$. Then $T_A=\Hom(A,A)$ is a complex with zero differential; an element $f\in(\fm\otimes T_A)^1$ satisfies the Maurer-Cartan equation iff $f^2=0$. For instance, put $R=k[\epsilon]/\epsilon^2$. Then any element $f$ of degree one is Maurer-Cartan. The corresponding to $f$ complex of $R$-modules is $R\otimes A$ as a graded $R$-module, and has $f$ as the differential. Suppose that all components of $f$ are non-zero. Then $(R\otimes A,f)$ is contractible and of course can not be thought of being a deformation of $A$. \section{Simplicial Deligne groupoid} \label{newnerve} \subsection{Definition} Let $k$ be a field of characteristic zero and $\fg\in\dgl(k)$ be a nilpotent dg Lie $k$-algebra. In this Section we construct a simplicial groupoid $\Gamma(\fg)=\{\Gamma_n(\fg)\}$ whose nerve (see~\ref{snerve}) is naturally homotopically equivalent to the nerve $\Sigma(\fg)$. The construction is a generalization (and a simplification) of the one we used in~\cite{uc}, 9.7.6. Recall (see~\cite{ddg},\cite{uc}, 8.1.1) that the nerve $\Sigma(\fg)$ of the nilpotent dg Lie algebra $\fg$ is defined as \begin{equation} \Sigma_n(\fg)=\MC(\Omega_n\otimes\fg), \label{nerve} \end{equation} $\Omega_n$ being the algebra of polynomial differential forms on the standard $n$-simplex. Following~\cite{uc}, Sect. 8, define a simplicial group $G=G(\fg)$ by the formula \begin{equation} G_n=\exp(\Omega_n\otimes\fg)^0. \label{G.} \end{equation} Here $\Omega_n\otimes\fg$ is a nilpotent dg Lie algebra, so its zero component is an honest nilpotent Lie algebra, and therefore its exponent makes sense. Define a simplicial groupoid $\Gamma:=\Gamma(\fg)$ (we will call it {\em simplicial Deligne groupoid} since its zero component is the conventional Deligne groupoid~\cite{gm}) as follows. $\Ob\Gamma=\MC(\fg);$ $ \Hom_{\Gamma}(x,y)_n=\{g\in G_n|g(x)=y\}.$ It is useful to have in mind the following easy \subsubsection{} \begin{lem}{G-contractible} The simplicial group $G(\fg)$ is always contractible. \end{lem} \begin{pf} As a simplicial set, $G$ is isomorphic to the simplicial vector space $$ n\mapsto (\Omega_n\otimes\fg)^0.$$ The latter is a direct sum of simplicial vector spaces of form $\Omega_{\bullet}^p$ (each one $\dim\fg^{-p}$ times) which are all contractible --- see~\cite{l}, p.~44. \end{pf} \subsection{Equivalence} Recall that any simplicial category (and more generally, any $\CC\in\Delta^0\Cat$) defines a bisimplicial set whose diagonal is called {\em the nerve} of $\CC$, denoted by $\CN(\CC)$ --- see~\ref{snerve}. \subsubsection{} \begin{prop}{eq.defns} The nerve $\Sigma(\fg)$ of a nilpotent dg Lie algebra is naturally homotopically equivalent to $\CN(\Gamma(\fg))$. \end{prop} \begin{pf}Define $\Gamma'\in\Delta^0\Grp$ (a simplicial groupoid in the wide sense) by the following formulas. $ \Ob\Gamma'_n=\MC(\Omega_n\otimes\fg);$ $ \Hom_{\Gamma'}(x,y)_n=\{g\in G_n(\fg)|g(x)=y\}.$ One has a natural fully faithful embedding $\Gamma(\fg)\to\Gamma'$. According to~\cite{uc}, 8.2.5, the map $\Gamma_n(\fg)\to\Gamma'_n$ is an equivalence of groupoids for each $n$. This implies that the induced map of the nerves $$\CN(\Gamma)\to\CN(\Gamma')$$ is a homotopy equivalence. Now we shall compare the nerve $\CN(\Gamma')$ to $\Sigma(\fg)$. Look at $\Gamma'$ as at a bisimplicial set. One has $$ \Gamma'_{pq}=\Sigma_p(\fg)\times G_p(\fg)^q.$$ This means that the simplicial set $\Gamma'_{\bullet q}$ is equal to $\Sigma(\fg)\times G(\fg)^q$. The simplicial set $G(\fg)$ is contractible by~\Lem{G-contractible}. Therefore, $\Gamma'_{\bullet q}$ is canonically homotopy equivalent to $\Sigma(\fg)$. This implies that the nerve $\CN(\Gamma')$ is homotopy equivalent to $\Sigma(\fg)$. \end{pf} \subsubsection{} \begin{rem}{gen}\Prop{eq.defns} generalizes the claim used in the proof of 9.7.6 of ~\cite{uc}. \end{rem} \subsubsection{} Let now $\fg\in\dgl(k)$. Following the well-known pattern, we define the functor $$\Gamma_{\fg}:\dgar(k)\to\sGrp$$ by the formula $$\Gamma_{\fg}(R)=\Gamma(\fm\otimes\fg)$$ for $(R,\fm)\in\dgar(k)$. The functor $\Gamma_{\fg}$ is also called the simplicial Deligne groupoid. \subsection{Properties} We wish to deduce now some properties of the simplicial Deligne groupoid functor which are similar to the properties of the nerve $\Sigma(\fg)$ --- see~\cite{uc}, Sect.~8. In what follows we use the closed model category (CMC) structure on the category $\sCat$ --- see~\ref{scat-cmc-ss}. \subsubsection{} \begin{prop}{gamma-fib} Let $f:\fg\to \fh$ be surjective (resp., a surjective quasi-isomorphism). Then for each $(R,\fm)\in\dgar(k)$ the map $$f:\Gamma_{\fg}(R)\to \Gamma_{\fh}(R)$$ is a fibration (resp., an acyclic fibration) in $\sCat$. \end{prop} \begin{pf} Note first of all that the similar claim holds for the nerve functor: according to~\cite{uc}, Prop. 7.2.1, the map $f:\Sigma_{\fg}(R)\to\Sigma_{\fh}(R)$ is a fibration (resp., acyclic fibration) provided $f$ is a surjection (resp., a surjective quasi-isomorphism). This implies that the map $$f:\Gamma_{\fg}(R)\to \Gamma_{\fh}(R)$$ satisfiest the property (1) of fibrations (resp., of acyclic fibrations) --- see~\ref{scat-fib},~\ref{scat-af}. Let us check the property (2). It claims that for any $x,y\in\Ob\Gamma_{\fg}(R)$ the map of simplicial sets $$ f:\uhom_{\fg}(x,y)\to\uhom_{\fh}(fx,fy)$$ is a Kan fibration (resp., acyclic Kan fibration) --- here we write $\uhom_{\fg}(\_,\_)$ instead of $\uhom_{\Gamma_{\fg}(R)}(\_,\_)$. Denote $G=G(\fg),\ H=G(\fh)$ the simplicial groups corresponding to $\fg$, $\fh$ as in the formula~\ref{G.}. A map from a simplicial set $K$ to $\uhom_{\fg}(x,y)$ is given by an element $g\in G(K)=\Hom(K,G)$ satisfying the condition $g(x)=y$. Let a commutative diagram in $\simpl$ \begin{center} $$\begin{CD} K@>>> \uhom_{\fg}(x,y) \\ @V{\alpha}VV @V{f}VV \\ L@>>> \uhom_{\fh}(fx,fy) \\ \end{CD}$$ \end{center} be given with $\alpha:K\to L$ being a cofibration of finite simplicial sets. We suppose also that either $\alpha$ or $f$ is a weak equivalence. Our aim is to find a map $L\to \uhom_{\fg}(x,y)$ commuting with the above diagram. Thus, we are given with a compatible pair of elements $g\in G(K),\ h\in H(L)$ satisfying the groperty $$ g(x)=y; \ h(fx)=fy.$$ Our aim is to lift this pair to an element $\widetilde{g}\in G(L)$ satisfying the property $\widetilde{g}(x)=y$. We will do this in two steps. First of all, since $f$ is surjective, the induced map of simplicial groups $f: G\to H$ is surjective, and, therefore, fibrant. Furthermore, since both $G$ and $H$ are contractible by~\Lem{G-contractible}, the map $f:G\to H$ is actually an acyclic fibration, and therefore the pair of compatible elements $g\in G(K),\ h\in H(L)$ lifts to an element $g'\in G(L)$. We can not, unfortunately, be sure that $g'(x)=y$. This is why we need the second step which will correct $g'$ to satisfy this property. Suppose $g'(x)=y'\in\MC(\Omega(L)\otimes\fm\otimes\fg)$. The elements $y$ and $y'$ of $\MC(\Omega(L)\otimes\fm\otimes\fg)$ have the same images in both $\MC(\Omega(K)\otimes\fm\otimes\fg)$ and $\MC(\Omega(L)\otimes\fm\otimes\fh)$. Now, the commutative diagram \begin{center} $$\begin{CD} \Omega(L)\otimes\fg@>>> \Omega(K)\otimes\fg\\ @VVV @VVV \\ \Omega(L)\otimes\fh@>>> \Omega(K)\otimes\fh\\ \end{CD}$$ \end{center} induces an acyclic fibration $$ p:\fg_1:=\Omega(L)\otimes\fg\to\Omega(K)\otimes\fg\times _{\Omega(K)\otimes\fh}\Omega(L)\otimes\fh=:\fg_2$$ of dg Lie algebras. Then the map $\Sigma_p:\Sigma_{\fg_1}(R)\to \Sigma_{\fg_2}(R)$ is an acyclic fibration. Now, we have two elements $y,y'\in\MC(\fm\otimes\fg_1)$ satisfying $p(y)=p(y')\in\MC(\fm\otimes\fg_2)$. Therefore, there exists an element $z\in \Sigma_{\fg_1}(R)_1$ such that $d_0z=y,d_1z=y'$ and $p(z)=s_0(p(y))$. Using the explicit description of $\Sigma_{\fg}(R)_1$ in~\cite{uc}, 8.2.3, one obtains and element $\gamma\in\exp(\fm\otimes\fg_1)$ satisfying $p(\gamma)=1\in\exp(\fm\otimes\fg_2);\ \gamma(y')=y$. Then one immediately sees that the element $\widetilde{g}=\gamma g'$ is the one we need. \end{pf} \subsubsection{} \begin{cor}{} 1. For any $\fg\in\dgl(k), R\in\dgar(k), x,y\in\Ob\Gamma_{\fg}(R)$ the simplicial set $\uhom(x,y)$ is fibrant. 2. Any quasi-isomorphism $f:\fg\to\fh$ induces a weak equivalence $$ f:\Gamma_{\fg}(R)\to\Gamma_{\fh}(R)$$ for each $R\in\dgar(k)$. \end{cor} \begin{pf} 1. Take $\fh=0$ in~\Prop{gamma-fib}. 2. The category $\dgl(k)$ admits a CMC structure with surjections as fibrations and quasi-isomorphisms as weak equivalences --- see~\cite{haha}, Sect.~4. Using this, present $f=p\circ i$ as a composition of an acyclic fibration $p$ and an acyclic cofibration $i$. Any acyclic cofibration in $\dgl(k)$ is left invertible: $q\circ i=\id$. The map $q$ is obviously an acyclic fibration. Then by~\Prop{gamma-fib} the map $f:\Gamma_{\fg}(R)\to\Gamma_{\fh}(R)$ is a weak equivalence. \end{pf} \section{Final} \label{final} \subsection{} We start with an observation explaining the connection between $T_A$ and the formal deformations of $A$. Let $B$ be a cofibrant $R\otimes\CO$-algebra with $(R,\fm)\in\dgar(k)$. Denote $A=k\otimes_RB$. The algebra $B$ is isomorphic, as a graded $\CO$-algebra, to $R\otimes A$. Choose a graded isomorphism $$ \theta: B\to R\otimes A$$ and put $$ z=\theta\circ d_B\circ\theta^{-1}-1\otimes d_A$$ where $d_B$ (resp., $d_A$) is the differential in $B$ (resp., in $A$). Then $z$ is a degree one derivation in $\fm\otimes T_A$ satisfying the Maurer-Cartan equation. A different choice of isomorphism $\theta$ gives rise to a Maurer-Cartan element $z'\in\fm\otimes T_A$ equivalent to $z$: there exists $g\in\exp(\fm\otimes T_A)^0$ such that $z'=g(z)$. In what follows we will use a (non-unique) presentation of a $R\otimes\CO$-algebra $B$ by an element $z\in\MC(\fm\otimes T_{k\otimes_RB})$. \subsection{Proof of the Theorem} To simplify the notation, denote $\CW=\CW^c_*(\CO,R),\ol{\CW}=\CW^c_*(\CO,k)$. \subsubsection{} \begin{lem}{isafibration} The natural map $\pi:W\to\ol{W}$ is a fibration in $\sCat$. \end{lem} \begin{pf} 1. Let us prove the condition (1) of Definition~\ref{scat-fib}. It means the following. Let $f:A\to B$ be a quasi-isomorphism of cofibrant $\CO$-algebras over $k$. Let one of two elements $a\in\MC(\fm\otimes T_A)$ or $b\in\MC(\fm\otimes T_B)$ be given. We have to check that there exists a choice of the second element and a map $$g: (R\otimes A,d+a)\to(R\otimes B,d+b)$$ of $R\otimes\CO$-algebras which lifts $f:A\to B$. We can consider separately the cases when $f$ is an acyclic fibration or an acyclic cofibration. In both cases we will be looking for the map $g$ in the form $$g=\gamma^{-1}_B\circ (\id_R\otimes f)\circ\gamma_A$$ where $\gamma_A\in\exp(\fm\otimes T_A)^0$ and similarly for $\gamma_B$. A map $g$ as above should commute with the differentials $d+a$ and $d+b$. This amounts to the condition $$ f_*(\gamma_A(a))=f^*(\gamma_B(b)),$$ where the natural maps $$ T_A\overset{f_*}{\lra}\Der_f(A,B)\overset{f^*}{\lla}T_B$$ are defined as in~\cite{haha}, 8.1. Recall that we are assuming that $f$ is either acyclic cofibration or an acyclic fibration. In both cases there exists a commutative square \begin{center} $$\begin{CD} T_f@>{\alpha}>> T_A \\ @V{\beta}VV @V{f_*}VV \\ T_B@>{f^*}>> \Der_f(A,B)\\ \end{CD}$$ \end{center} where $T_f$ is a dg Lie algebra and $\alpha,\beta$ are Lie algebra quasi-isomorphisms --- see~\cite{haha}, 8.2, 8.3. The maps $\alpha,\beta$ induce bijections $$ \pi_0(\Sigma_{T_A}(R))\lla\pi_0(\Sigma_{T_f}(R)) \lra\pi_0(\Sigma_{T_A}(R))$$ which prove the assertion. 2. Let us check the condition (2) of~\ref{scat-fib}. Let $\widetilde{A},\widetilde{B}\in\CW$ and let $A=k\otimes_R\widetilde{A}, B=k\otimes_R\widetilde{B}$. We have to check that the map \begin{equation} \uhom(\widetilde{A},\widetilde{B})\to\uhom(A,B) \label{red} \end{equation} is a Kan fibration. But this results from the SCMC structure on $\Alg(R\otimes\CO)$. In fact, $\widetilde{A}$ is cofibrant, and the reduction map $\widetilde{B}\to B$ can be considered as a fibration in $\Alg(R\otimes\CO)$. Therefore, the map \begin{equation} \uhom(\widetilde{A},\widetilde{B})\to\uhom(\widetilde{A},B) \label{red2} \end{equation} is a Kan fibration. But the maps~(\ref{red2}) and (\ref{red}) coincide, so the condition (2) of~\ref{scat-fib} is proven. \end{pf} \subsubsection{} Fix now a cofibrant $\CO$-algebra $A$ and denote $\fg=T_A$. Fix $(R,\fm)\in\dgar(k)$. Define a map of simplicial categories $$ \alpha:\Gamma_{\fg}(R)\to W$$ as follows. Let $z\in\MC(\fm\otimes\fg)=\Ob\Gamma_{\fg}(R)$. Put $$\alpha(z)=(R\otimes A, 1\otimes d+z).$$ Now, any element $g\in G_n=\exp(\Omega_n\otimes\fm\otimes\fg)^0$ defines a graded automorphism of $\Omega_n\otimes R\otimes A$. This obviously defines an isomorphism of $R\otimes\CO$-algebras $$ (R\otimes A, 1\otimes d+z)\lra (R\otimes A, 1\otimes d+g(z)).$$ The map $\alpha:\Gamma_{\fg}(R)\to W$ identifies $\Gamma_{\fg}(R)$ with the fibre of $\pi:W\to\ol{W}$ at $A$. Since $\pi$ is a fibration, this is weakly equivalent to the homotopy fibre of $\pi$. Now Theorem follows since the nerve functor $\CN$ preserves fibrations and weak equivalences by~\ref{nerve-exactness}. \subsection{Concluding remarks} \subsubsection{} The formula~(\ref{def-def-functor}) defines a deformation functor for operad algebras not necessarily concentrated in non-positive degrees. The definition seems to be correct also in this case, in spite of the fact that $T_A$ does not govern deformations of $A$ in this case. It seems that one should find in this case a more clever way to define the tangent Lie algebra. It might consist of ``tame'' derivations of a cofibrant resolution $\widetilde{A}$ of $A$, the ones which behave well with respect to a filtration on $\widetilde{A}$. \subsubsection{} One can define ``hard'' formal deformations of a $\CO$-algebra $A$ which deform not only the algebra $A$ itself but also the base operad $\CO$. Then the universal deformation of $(\CO,A)$ would provide the tangent Lie algebra $T_A$ with a canonical extra structure. \section{Appendix: simplicial categories} In this Section we present a more or less standard information about simplicial categories. It includes the description~\ref{scat-cmc} of a CMC structure on the category $\sCat$ of small simplicial categories. This structure is a slight generalization of the one described in~\cite{dk}. \subsection{Weak equivalences and fibrations in $\sCat$ } \label{scat-cmc-ss} Here we define a closed model category structure on the category $\sCat$ of simplicial categories. \subsubsection{(Co)limits} \label{lim} The category $\sCat$ admits arbitrary limits and colimits. Inverse limits in $\sCat$ are induced by inverse limits in $\Ens$ in the obvious sense. The existence of inductive limits in $\sCat$ follows by a general abstract nonsense from the existence of inductive limits in $\Ens$. Note that the functor $\sCat\to\Ens$ assigning to each simplicial category the set of its objects, commutes with inductive limits. The set of morphisms of an inductive limit is freely generated by the morphisms of all categories involved, modulo an obvious equivalence relation. Note that the existence of direct limits in $\sCat$ allows one to mimic the procedure of ``adding variables''. We will single out the following cases. {\em Adding an object.} Given $\CC\in\sCat$, denote $\CC\langle *\rangle$ the coproduct of $\CC$ with the trivial one-object category $*$. {\em Adding an ingoing arrow.} Given $\CC\in\sCat,\ x\in\Ob\CC$, one defines $\CC\langle *\to x\rangle$ with the set of objects $\Ob\CC\coprod\{*\}$ and the set of morphisms freely generated by $\Mor\CC$ and by the map $*\to x$. {\em Adding an outgoing arrow.} The category $\CC\langle x\to *\rangle$ is defined similarly to the above. {\em Adding maps between objects} Given $\CC\in\sCat$, $x,y\in\Ob\CC$ and a map $\alpha:\uhom_{\CC}(x,y)\to H$ of simplicial sets, the simplicial category $\CC\langle x,y;\alpha\rangle$ has the same objects as $\CC$. Its set of morphisms is freely generated by $\Mor\CC$ and by $H$. \subsubsection{} Define the functor $$\pi_0:\sCat\to\Cat$$ as follows. For $\CC\in\sCat$ the category $\pi_0(\CC)$ has the same objects as $\CC$. For $x,y\in\Ob\pi_0(\CC)$ $$\Hom_{\pi_0(\CC)}(x,y)=\pi_0(\Hom_{\CC}(x,y)).$$ \subsubsection{} \begin{defn}{scat-we} A map $f:\CC\to\CDD$ in $\sCat$ is called a weak equivalence if the following properties are satisfied. (1) The map $\CN(\pi_0(f))$ is a weak equivalence of simplicial sets. (2) For all $x,x'\in\Ob\CC$ the map $f:\uhom(x,x')\to\uhom(fx,fx')$ is a weak equivalence. \end{defn} \subsubsection{} \begin{defn}{scat-fib}A map $f:\CC\to\CDD$ in $\sCat$ is called a fibration if it satisfies the following properties (1) the right lifting property (RLP) with respect to ``adding an ingoing or an outgoing arrow'' $$\CA\to\CA\langle *\to x\rangle,\ \CA\to\CA\langle x\to *\rangle$$ (see~\ref{lim}). (2) For all $x,x'\in\Ob\CC$ the map $f:\uhom(x,x')\to\uhom(fx,fx')$ is a Kan fibration. This is equivalent the the RLP with respect to all maps $\CA\to\CA\langle x,y;\alpha\rangle$ where $\alpha$ is an acyclic fibration (see~\ref{lim}). \end{defn} \subsubsection{} \begin{thm}{scat-cmc}The category $\sCat$ admits a CMC structure with weak equivalences described in~\ref{scat-we} and fibrations as in~\ref{scat-fib}. \end{thm} \subsubsection{} An explicit description of different classes of morphisms in $\sCat$ is given below. The proof of the Theorem is standard. It is given in~\ref{pf-scat-cmc}. \subsubsection{} A map $f:\CC\to\CDD$ in $\sCat$ is called an acyclic fibration if it is simultaneously a weak equivalence and a fibration. \begin{lem}{scat-af} A map $f:\CC\to\CDD$ is an acyclic fibration iff the following conditions are satisfied. (1) the map $\Ob f:\Ob\CC\to\Ob\CDD$ is surjective. In other words, $f$ satisfies the RLP with respect to ``adding an object map'' $\CA\to\CA\langle *\rangle$. (2) For all $x,x'\in\Ob\CC$ the map $f:\uhom(x,x')\to\uhom(fx,fx')$ is an acyclic Kan fibration. \end{lem} \begin{pf}If $f$ satisfies (1), (2), it is clearly an acyclic fibration. In the other direction, suppose $f$ is an acyclic fibration. Then the property (2) is clear. We have only to check that $\Ob f$ is surjective. Since $f$ satisfies the RLP with respect to ingoing and outgoing arrows, $\CDD$ is a disjoint union of the full subcategories, defined by the image of $\Ob f$ and by its complement. Since $\pi_0(f)$ is a weak equivalence, it induces a bijection of the connected components of $\CC$ and $\CDD$ and this proves the claim. \end{pf} \subsubsection{} \label{sc} A map $f:\CC\to\CDD$ will be called {\em a standard cofibration} if there is a collection of maps $f_i:\CC_i\to\CC_{i+1},\ i\in\Bbb{N}$ such that $\CC=\CC_0,\ \CDD=\dirlim\CC_i$, and each $f_i$ is a coproduct of maps of one of the following two types: (1) Adding an object $\CC_i\to\CC_i\langle *\rangle$; (2) Adding maps between objects $\CC_i\to\CC_i\langle x,y;\alpha\rangle$ with $\alpha$ injective. By~\ref{scat-af}, standard cofibrations satisfy the LLP with respect to all acyclic fibrations. \subsubsection{} \label{sac} A map $f:\CC\to\CDD$ will be called {\em a standard acyclic cofibration} if there is a collection of maps $f_i:\CC_i\to\CC_{i+1},\ i\in\Bbb{N}$ such that $\CC=\CC_0,\ \CDD=\dirlim\CC_i$, and each $f_i$ is a coproduct of maps of one of the following three types: (1+) Adding an ingoing arrow $\CC_i\to\CC_i\langle *\to x\rangle$; (1--) Adding an outgoing arrow $\CC_i\to\CC_i\langle x\to *\rangle$; (2) Adding maps between objects $\CC_i\to\CC_i\langle x,y;\alpha\rangle$ with $\alpha$ acyclic cofibration. By~\ref{scat-fib}, standard acyclic cofibrations satisfy the LLP with respect to all fibrations. \subsubsection{} The following description of cofibrations and of acyclic cofibrations results from the proof of~\Thm{scat-cmc}. \begin{cor}{c-and-ac}1. Any cofibration in $\sCat$ is a retract of a standard cofibration. 2. Any acyclic cofibration in $\sCat$ is a retract of a standard acyclic cofibration. \end{cor} \subsection{Proof of~\Thm{scat-cmc}} \label{pf-scat-cmc} The axioms (CM 1), (CM 2), (CM 3), (CM 4)(ii) are immediately verified. (CM 5)(ii) Let $f:X\to Y$ be a map in $\sCat$. Adding objects to $X$, we can ensure that the map $f:\Ob(X)\to\Ob(Y)$ is surjective. Then, adding maps between objects, we can decompose $f$ into a standard cofibration followed by an acyclic fibration. This implies, in particular, that any cofibration is a retract of a standard cofibration. To check the axiom (CM 5)(i) we need the following \subsubsection{} \begin{lem}{sac-is-ac} Standard acyclic cofibrations are acyclic cofibrations. \end{lem} \begin{pf}It is enough to prove that a map $\CC\to\CDD$ is a weak equivalence when $\CDD$ is obtained from $\CC$ by one of the following ways. (1) adding a number of ingoing arrows; (2) adding a number of outgoing arrows; (3) adding (simultaneously) maps between objects $x_i$ and $y_i$ along acyclic cofibrations $\alpha_i:\uhom(x_i,y_i)\to H_i$. In the first two cases the map $\CC\to \CDD$ is easily split by an acyclic fibration. The shortest way to get the result in the case (3) is to use Proposition 7.2 of~\cite{dk} which claims the existence of CMC structure on the category of simplicial categories having a fixed set of objects. \end{pf} (CM 5)(i) Let $f:X\to Y$ be a map in $\sCat$. Adding ingoing and outgoing arrows to $X$, we can ensure that the image of $\Ob(X)$ under $f$ consists of a number of connected components of $\Ob(Y)$. From now on we can suppose, without loss of generality, that $f$ is surjective on objects. Then for a decomposition $f=p\circ i$ it is enough to check that $p$ satisfies condition (2) of ~\ref{scat-fib} to ensure $p$ is a fibration. Now applying step by step the procedure of adding maps between objects $\CC\to\CC\langle x,y;\alpha\rangle$ along acyclic cofibrations $\alpha$, we can construct a decomposition $f=p\circ i$ with $p$ fibration and $i$ a standard acyclic cofibration. According to~\Lem{sac-is-ac}, $i$ is an acyclic cofibration. Now, applying the proof of (CM 5)(i) to any acyclic cofibration $f$, we deduce that $f$ is a retract of a standard acyclic cofibration. (CM 4)(i) By definition, any standard acyclic cofibration satisfies LLP with respect to all fibrations. Any acyclic fibration is a retract of a standard acyclic fibration, and therefore satisfies as well LLP with respect to all fibrations. Theorem is proven. \subsection{Simplicial nerve} \label{snerve} \subsubsection{} \label{snerve} In what follows we identify $\Cat$ with the full subcategory of $\simpl$. Then every simplicial category (and even every $\CC\in\Delta^0\Cat$) can be seen as a bisimplicial set; its diagonal will be called {\em the nerve} of $\CC$ and will be denoted $\CN(\CC)$. If $\CC$ is a ``usual'' category, $\CN(\CC)$ is its ``usual'' nerve. The functor $\CN:\sCat\to\simpl$ admits a left adjoint functor $$\SCAT:\simpl\to\sCat$$ defined by the properties \begin{itemize} \item{$\Ob\SCAT(\Delta^n)=[n]=\{0,\ldots,n\};$} \item{$\Mor\SCAT(\Delta^n)\text{ is freely generated by }a_i\in\uhom_n(i-1,i), \ i=1,\ldots,n;$} \item{ $\SCAT$ commutes with arbitrary colimits.} \end{itemize} \subsubsection{} \begin{prop}{nerve-exactness} The nerve functor $\CN:\sCat\to\simpl$ preserves weak equivalences, fibrations and cofibrations. \end{prop} \begin{pf} 1. To check that $\CN$ preserves the fibrations, it is enough to prove that the adjoint functor $\SCAT$ preserves acyclic cofibrations. For this we have to check that $\SCAT$ transforms any map $\Lambda^n_i\to\Delta^n$ to an acyclic fibration. This is an easy exercise (one should consider the cases $n=1$ and $n>1$ separately). Note that the same reasoning (even easier!) proves that $\CN$ preserves acyclic fibrations --- this is because $\SCAT$ preserves cofibrations. 2. It is clear that $\CN(f)$ is a weak equivalence provided $f$ is a weak equivalence {\em bijective on objects}. To prove the general claim, we present $f$ as a composition of an acyclic fibration with an acyclic cofibration and therefore reduce the problem to the case $f$ is an acyclic cofibration. Using~\ref{c-and-ac}, we can suppose that $f$ is of one of the types (1+), (1--), (2) of~\ref{sac}. The type (2) does not change the set of objects, so we have nothing to prove. The maps of types (1+), (1--) split, and the splitting map is an acyclic fibration. This proves the claim. 3. The claim about cofibrations is obvious. \end{pf}
\section{INTRODUCTION} We make use of recently proposed parameterizations \cite{boya} (cf. \cite{zhsl}) --- in terms of squared components of the unit $(n-1)$-sphere and the $n \times n$ unitary matrices --- of the $n \times n$ density matrices. First (in sec.~\ref{buressection}), we derive (prior) probability distributions of particular interest over both the three-dimensional convex set of two-state quantum systems and the eight-dimensional convex set of three-state quantum systems. These distributions are the normalized volume elements of the corresponding Bures metrics on these systems. Hall \cite{hall} (cf. \cite{hall2,stasz,rieffel}) has contended that such distributions correspond to ``minimal-knowledge'' ensembles, that is the most {\it random} ensembles of possible states. In particular, for the two-dimensional quantum systems, he argues that the Bures metric provides such a minimal-knowledge ensemble, since it ``corresponds to the surface of a unit four-ball, i. e., to the maximally symmetric space of positive curvature \ldots This space is homogeneous and isotropic, and hence the Bures metric does not distinguish a preferred location or direction in the space of density operators'' \cite{hall}. Somewhat contrarily though, Slater \cite{slaterpla1} has reported results (based on the concept of comparative noninformativities of priors, first expounded in \cite{clarke}) that indicate the Bures metric generates ensembles that are less noninformative than other (monotone) metrics of interest. The Bures metric fulfills the role of the {\it minimal} monotone metric \cite{petzsudar,petzlaa,lesniewski}, and has been the focus of a considerable number of studies \cite{hall,slaterpla1,hubner1,hubner2,brauncaves,dittmann1,dittmann2}. ``An infinitesimal statistical distance has to be monotone under stochastic mappings'' \cite[p. 786]{petzsudar}. All stochastically monotone Riemannian metrics are characterized by means of operator monotone functions. Among all (suitably normalized) operator montone functions $f(t)$ with $f(1)=1$ and $f(t) = t f(t^{-1)}$, there is a minimal and a maximal one \cite{kubo}. (The concept of a minimal metric was apparently introduced by Zolotarev in his extensive paper, ``Metric distances in spaces of random variables and of their distributions'' \cite[sec. 1.4]{zolotarev}, but it is not entirely clear that the meaning there is the same as in the terminology ``minimal monotone metric''.) We should bear in mind, though --- as emphasized by Petz and Sud\'ar \cite{petzsudar} --- that, in strong contrast to the classical situation, in the quantum domain there is not a unique monotone metric, but rather a (nondenumerable) multiplicity of them. Their comparative properties need to be evaluated, before deciding which specific one to employ for a particular application. We repeat the concluding remarks of Petz and Sud\'ar: ``Therefore, more than one privileged metric shows up in quantum mechanics. The exact clarification of this point requires and is worth further studies.'' We have previously reported \cite{slaterjmp} (in terms of parameterizations other than that of Boya {\it et al} \cite{boya}) the Bures probability distribution for the two-state systems, and also \cite{slaterjpa} for an imbedding of these systems into a four-dimensional convex set of three-state systems, but the result below (\ref{new}) for the full eight-dimensional convex set of three-state systems is clearly novel in nature. In fact, in \cite[sec. II.E]{slatertherm2} we discussed certain (unsuccessful) efforts in these directions (although the volume element of the {\it maximal} monotone metric --- which is not strictly normalizable --- proved more amenable to analysis there). In sec.~\ref{bureshallsection}, we determine certain necessary elements for extending the work reported in sec.~\ref{buressection} to the higher-dimensional quantum systems ($n >3$). This involves finding the normalization constant ($C_{n}$), explicitly first discussed by Hall \cite[eq. (25)]{hall}, for the {\it marginal} Bures prior probability distribution over the $(n-1)$-dimensional simplex of the $n$ eigenvalues of the $n \times n$ density matrices. These constants are found to exhibit quite remarkable number-theoretic properties. It would, therefore, certainly be of substantial interest to find a general formula for $C_{n}$. Knowledge of the value of $C_{n}$, together with that of the invariant Haar element for $SU(n)$ --- apparently presently available, however, in suitably parameterized form (cf. \cite{Zycz1,Zycz2}) for $n \leq 3$ \cite{byrd,byrdsudarshan} --- would allow one to construct the Bures prior probability distribution itself for the $n$-level quantum systems. In an extensive study \cite{kratt}, Krattenthaler and Slater examined (in the framework of the {\it two}-state systems) the hypothesis that the normalized volume element of the Bures metric would function in the quantum domain in a role parallel to that fulfilled classically by the ``Jeffreys' prior'' --- that is, the normalized volume element of the unique monotone/Fisher information metric \cite{kass,kwek}. In particular, they were interested in \cite{kratt} in the possibility of extending certain (classical) seminal results of Clarke and Barron \cite{cb1,cb2}. They did conclude, however, contrary to their working hypothesis, that the normalized volume element of the Bures metric does not in fact strictly fullfill the same role as the Jeffreys' prior (in yielding both the asymptotic minimax and maximin redundancies for universal coding/data compression), but it appears to come remarkably close to doing so (cf. Fig.~\ref{comparison}). In sec.~\ref{QUASI}, for the cases $n=2$ and 3, the ``quasi-Bures'' prior probability distributions are presented that appear to fulfill this distinguished information-theoretic role. \section{BURES PROBABILITY DISTRIBUTIONS OVER THE ${n \times n}$ DENSITY MATRICES} \label{buressection} Boya {\it et al} \cite{boya} have recently ``shown that the mixed state density matrices for $n$-state systems can be parameterized in terms of squared components of an $(n-1)$-sphere and unitary matrices''. The mixed state density matrix ($\rho$) is represented in the form, \begin{equation} \label{one} \rho = U D U^{\dagger} , \end{equation} where $U$ denotes an $SU(n)$ matrix, $U^{\dagger}$ its conjugate transpose and $D$ a diagonal density operator, the diagonal entries ($d_{i}$'s) of which --- being the eigenvalues of $\rho$ --- are the squared components of the $(n-1)$-sphere. Thus, for $n =2$, \begin{equation} \label{two} D = \pmatrix{\cos^{2}{\theta /2} & 0 \cr 0 & \sin^{2}{\theta /2} \cr} \qquad 0 \leq \theta \leq {\pi \over 2}, \end{equation} and, for $n = 3$, \begin{equation} \label{three} D = \pmatrix{\cos^{2}{\phi /2} \sin^{2}{\theta /2} & 0 & 0 \cr 0 & \sin^{2}{\phi /2} \sin^{2}{\theta /2} & 0 \cr 0 & 0 & \cos^{2}{\theta /2} \cr} \qquad 0 \leq \theta, \phi \leq \pi. \end{equation} (Note the differences in the ranges of angles used in the two cases. This will be commented upon in sec.~\ref{n4bures}.) Biedenharn and Louck have presented \cite[eq. (2.40)]{bied} the parameterization of an element of $SU(2)$, \begin{equation} \label{su2} U(\alpha \beta \gamma) = e^{-i \alpha \sigma_{3} /2} e^{-i \beta \sigma_{2} /2} e^{-i \gamma \sigma_{3} /2}, \end{equation} in terms of the Pauli matrices ($\sigma_{i}$'s) and three Euler angles --- $ 0 \leq \alpha < 2 \pi, 0 \leq \beta \leq \pi, 0 \leq \gamma < 2 \pi$ --- with an associated invariant Haar measure \cite[eq. (3.134)]{bied}, \begin{equation} \label{invariant1} \mbox{d} \Omega_{2} = {1 \over 8} \mbox{d} \alpha \mbox{d} \gamma \sin{\beta} \mbox{d} \beta. \end{equation} Byrd \cite{byrd} (cf. \cite{byrdsudarshan}) has extended this approach to $SU(3)$. He obtains \begin{equation} U(\alpha,\beta,\gamma,\kappa,a,b,c,\zeta) = e^{i \lambda_{3} \alpha} e^{i \lambda_{2} \beta} e^{i \lambda_{3} \gamma} e^{i \lambda_{5} \kappa} e^{i \lambda_{3} a} e^{i \lambda_{2} b} e^{i \lambda_{3} c} e^{i \lambda_{8} \zeta}, \end{equation} where $\lambda_{i}$ denotes one of the eight $3 \times 3$ Gell-Mann matrices \cite{lukach}. The corresponding invariant element is \begin{equation} \label{invariant2} \mbox{d} \Omega_{3} = \sin{2 \beta} \sin{2 b} \sin{2 \kappa} {\sin^{2}{\kappa}} \mbox{d} \alpha \mbox{d} \beta \mbox{d} \gamma \mbox{d} \kappa \mbox{d} a \mbox{d} b \mbox{d} c \mbox{d} \zeta, \end{equation} with the eight Euler angles having the ranges, \begin{equation} \label{ranges} 0 \leq \alpha, \gamma, a, c < \pi, \quad 0 \leq \beta, b, \kappa \leq {\pi /2}, \quad 0 \leq \zeta < \sqrt{3}. \end{equation} For our purposes, the Euler angle $\gamma$ for the case $n=2$ and $c$ and $\zeta$ in the case $n=3$ are irrelevant, as they ``drop out'' in the formation of the product (\ref{one}). (I thank M. Byrd for this important observation.) So, we will employ below the appropriate {\it conditional} versions of these invariant measures (\ref{invariant1}) and (\ref{invariant2}) --- the condition (a technical statistical term) corresponding, of course, to the ignoring of the indicated angles. The Bures metric itself is expressible in the form \cite[eq. (10)]{hubner1} \begin{equation} \label{bures} d_{B}^{2}(\rho,\rho + \mbox{d} \rho) = \sum_{i,j =1}^{n} {1 \over 2} {{|<i|\mbox{d} \rho |j>}^{2} \over d_{i} + d_{j}}, \end{equation} where $|i>$ denotes the eigenvectors of the $n \times n$ density matrix $\rho$, $<j|$, the corresponding complex conjugate (dual) vectors, and the $d$'s are the associated eigenvalues. The parameterization of Boya {\it et al} is, then, particularly convenient, since the eigenvalues and eigenvectors of $\rho$ are immediately available. Our chief concern must, then, be to compute the complete Jacobian of the transformation to the set of parameters of Boya {\it et al}. (We shall note for further reference the occurrence of the term $d_{i} + d_{j}$ in (\ref{bures}). This, is of course, simply proportional to the arithmetic mean, $(d_{i} +d_{j})/2$. By replacing this term by (twice) the {\it exponential/identric} mean (\ref{qi13}) of $d_{i}$ and $d_{j}$, that is $2 I(d_{i},d_{j})$, we shall obtain the particular ``quasi-Bures'' distributions described in sec.~\ref{QUASI}.) \subsection{The Bures case $n=2$} For the case $n=2$, the volume element of the Bures metric (\ref{bures}) is proportional to the product of the inverse of the square root of the determinant of $\rho$ (or, equivalently, the determinant of $D$) with two Jacobians. The first Jacobian (in line with the familiar practice in the theory of random matrices \cite[eq. (3.3.5)]{mehta}) is itself the product of $(d_{1}-d_{2})^2$ and the (conditional) invariant element (\ref{invariant1}). The second Jacobian, ${\sin{\theta} / 2}$, corresponds simply to the transformation from cartesian coordinates to the squared polar coordinates employed in (\ref{two}). Simplifying and normalizing the full product, we arrive at the probability density for the normalized volume element of the Bures metric over the three-dimensional convex set of two-state quantum systems. This density is \begin{equation} \label{n2} \mbox{d} p_{Bures:2}(\theta,\alpha,\beta) = {\cos^{2}{\theta} \sin{\beta} \over {\pi^{2}}} \mbox{d} \theta \mbox{d} \alpha \mbox{d} \beta \qquad 0 \leq \theta \leq {\pi \over 2}, \quad 0 \leq \alpha \leq 2 \pi, \quad 0 \leq \beta \leq \pi. \end{equation} The expected values of the eigenvalues are, then, ${1/2 \pm 4 / 3 \pi}$. \subsection{The Bures case $n=3$} \label{n3bures} For the case $n=3$, the volume element of the Bures metric is equal to the product of: (i) two Jacobians again, one of which now has the form $\big((d_{1}- d_{2}) (d_{1}-d_{3}) (d_{2}-d_{3})\big)^2$ multiplied by the (conditional) invariant measure (\ref{invariant2}), while the other, $(\cos{{\theta \over 2}} \sin^{3}{{\theta \over 2}} \sin{\phi})/2$, corresponds to the transformation to squared spherical coordinates used in (\ref{three}); and (ii) the reciprocal of the product of the square root of the determinant of $\rho$ (or, equivalently, of $D$) and the difference between the sum of the three principal minors of order two of $\rho$ (or, equivalently, of $D$) and the determinant itself. Since $|\rho|^{1/2} = (\cos{{\theta \over 2}} \sin^{2}{{\theta \over 2}} \sin{\phi})/2$, it can be seen that considerable cancellation occurs between the numerator and the denominator of the full product. The normalization of the resultant volume element required considerable manipulations using MATHEMATICA (basically involving reducing the problem to the simplest possible form at each stage of the integration process). We obtained the following Bures prior probability density over the eight-dimensional convex set of three-state (spin-1) quantum systems, \begin{equation} \label{new} \mbox{d} p_{Bures:3}(\theta,\phi,\alpha,\beta,\gamma,\kappa,a,b) = {35 u \over 128 {\pi}^4 (35 + 28 \cos{\theta} + \cos{2 \theta} -8 \cos{2 \phi} {\sin^{4}{{\theta \over 2}}})} \mbox{d} \theta \mbox{d} \phi \mbox{d} \widetilde{\Omega}_{3}, \end{equation} where \begin{equation} \label{newvariable} u = {\sin^{3}{{\theta \over 2}}} \big((35 + 60 \cos{\theta} + 33 \cos{2 \theta} ) \cos{\phi} - {8 \cos{3 \phi}} {{\sin^{4}{{\theta \over 2}}}})^2 \big), \end{equation} and the conditional invariant element (cf. (\ref{invariant2})) is \begin{equation} \label{conditionalhaar} \mbox{d} \widetilde{\Omega}_{3} = \sin{2 \beta} \sin{2 b} \sin{2 \kappa} {\sin^{2}{\kappa}} \mbox{d} \alpha \mbox{d} \beta \mbox{d} \gamma \mbox{d} \kappa \mbox{d} a \mbox{d} b. \end{equation} The eight variables have the previously indicated ranges ((\ref{three}), (\ref{ranges})). In Fig.~\ref{twodim}, we display the two-dimensional marginal probability distribution of (\ref{new}) over the parameters $\theta$ and $\phi$ (which are invariant under unitary transformations of $\rho$). \begin{figure} \centerline{\psfig{figure=buresprior2.eps}} \caption{Bivariate marginal Bures prior probability distribution over the variables $\theta$ and $\phi$ for three-state systems} \label{twodim} \end{figure} Let us note that the fully mixed state --- corresponding to the $3 \times 3$ diagonal density matrix with entries equal to $1/3$ --- is obtained at $\phi = \pi/2, \theta = 2 \cos^{-1}({1/\sqrt{3})} \approx 1.91063$. The probability density (\ref{new}) is zero at this distinguished point, as well as along the loci $\theta = 0$ and $\phi = \pi /2$. (Wherever at least two of the eigenvalues of $\rho$ or, equivalently $D$, are equal, the density is zero.) The one-dimensional marginal probability density (Fig.~\ref{onedim1}), obtained by integrating (\ref{new}) over all variables except $\theta$, is \begin{equation} \mbox{d} \tilde{p}_{Bures:3}(\theta) = {35 \over 256} (-1533 + 2816 \cos{{\theta \over 2}} - 1988 \cos{\theta} + 1152 \cos{{3 \theta \over 2}} - 447 \cos{2 \theta} + 128 \cos{{5 \theta \over 2}}) \sin^{3}{{\theta \over 2}} \mbox{d} \theta. \end{equation} \begin{figure} \centerline{\psfig{figure=buresprior1.eps}} \caption{Univariate marginal Bures prior probability distribution over the variable $\theta$ for three-state systems} \label{onedim1} \end{figure} The relative maxima of this density are located at .914793, 2.2795 and $\pi$, while the relative minima are at 0, 1.59995 and 2.61732. The one-dimensional marginal probability density (Fig.~\ref{onedim2}), obtained by integrating (\ref{new}) over all variables except $\phi$, is \begin{equation} \label{othermarginal} \mbox{d} \tilde{\tilde{p}}_{Bures:3}(\phi) = {1 \over 768 \pi} \big(\cot{\phi} {\csc^{8}{\phi}} \big(110100480 \tan^{-1}{[\cot{\phi \over 2}]} {\cos^{12}{{\phi \over 2}}} - 26880 (792 (2 \pi -\phi) \cos{\phi} + 8 \pi \end{equation} \begin{displaymath} (55 \cos{3 \phi} + 3 \cos{5 \phi}) + \phi (495 \cos{2 \phi} - 220 \cos{3 \phi} +66 \cos{4 \phi} -12 \cos{5 \phi} + \cos{6 \phi})) + 16885656 \sin{2 \phi} \end{displaymath} \begin{displaymath} + 5069937 \sin{4 \phi} + 167012 \sin{6 \phi} - 3 ( 4139520 \phi + 124 \sin{8 \phi} - 4 \sin{10 \phi} + \sin{12 \phi})\big)\big) \mbox{d} \phi. \end{displaymath} \begin{figure} \centerline{\psfig{figure=buresprior1other.eps}} \caption{Univariate marginal Bures prior probability distribution over the variable $\phi$ for three-state systems} \label{onedim2} \end{figure} The limits of (\ref{othermarginal}) as $\phi$ approaches 0 and $\pi$ are both equal to ${20 \over 9 \pi} \approx .707355$. (The density is highly oscillatory in the vicinity of these boundary points.) MATHEMATICA does, in fact, perform a {\it symbolic/exact} integration of (\ref{othermarginal}), yielding the result 1. (Note that for our analyses below for $n>3$, we have found it necessary to rely upon {\it numerical} integrations, although the results obtained do appear to indicate that exact solutions exist, which in principle might be found with a powerful enough computer.) However, several warning messages are generated en route to this result, concerning indeterminate expressions and inconsistencies in the arguments of MeijerG functions (which are very general forms of hypergeometric functions) \cite{MeijerG}. \section{HALL NORMALIZATION CONSTANTS FOR {\it MARGINAL} BURES PROBABILITY DISTRIBUTIONS OVER THE $(n-1)$-DIMENSIONAL SIMPLEX OF THE $n$ EIGENVALUES OF THE $n \times n$ DENSITY MATRICES} \label{bureshallsection} Let us note that Hall \cite[eq. (24)]{hall} has, in fact, given an explicit formula for the volume element of the Bures metric on the $n \times n$ density matrices. This is (converting to the notation used above), \begin{equation} \label{hallformula} \mbox{d} V_{Bures:n} = {\mbox{d} d_{1} \ldots \mbox{d} d_{n} \over (d_{1} \ldots d_{n})^{1/2}} \prod_{i < j}^{n} 4 {(d_{j} -d_{k})^{2} \over d_{i} + d_{j}} \mbox{d} x_{ij} \mbox{d} y_{ij}, \end{equation} where the real part of the $ij$-entry of the diagonalizing unitary matrix $U$ in formula (\ref{one}) is represented by $x_{ij}$ and the imaginary part by $\pm i y_{ij}$. Since in the parameterization of Boya {\it et al} \cite{boya}, which we have employed, one uses not these $x$'s and $y$'s, but rather the Euler angles parameterizing the unitary matrix $U$, we have been compelled to replace the differential elements $\mbox{d} x_{ij} \mbox{d} y_{ij}$ in (\ref{hallformula}) by the corresponding conditional form ($\mbox{d} \widetilde{\Omega}$) of the invariant (Haar) measures (\ref{invariant1}) and (\ref{invariant2}). One can, then, confirm that our presentation and results are fully consistent with the use of (\ref{hallformula}), bearing in mind the unit trace requirement that $d_{1} + \ldots +d_{n} =1$. Hall \cite[eq. (25)]{hall} also expressed the marginal Bures probability distribution over the space of $n$ eigenvalues of $\rho$ as \begin{equation} \label{hallconstant} \tilde{p}_{Bures:n}(d_{1},\ldots,d_{n}) = C_{n} {\delta (d_{1} + \ldots + d_{n} -1) \over (d_{1} \ldots d_{n})^{1/2}} \prod_{i < j}^{n} {(d_{i} -d_{j})^2 \over d_{i}+ d_{j}}. \end{equation} We shall report the values of the ``Hall constants'' $C_{n}$ for $n =2, 3, 4$ and 5, immediately below. (The results for $n > 2$ are, apparently, new.) \subsection{The Hall constants for $n=2$ and $n=3$} \label{proximate} If, for consistency with our further results for $n>2$, we take $\theta \in [0,\pi]$ and not $[0,\pi /2]$ as in \cite{boya}, then we find that $C_{2}$ equals $2 / \pi \approx .63662$ (cf. \cite[eq. (30)]{hall}). From the results of the analysis in sec.~\ref{n3bures}, we are able to determine, for the first time, apparently, that $C_{3} = 35 / \pi \approx 11.1408$. Let us note here that 35 is, of course, simply the product of {\it proximate} or {\it neighboring} prime numbers, that is, $35 = 5 \cdot 7$. \subsection{The Hall constant for $n=4$ and associated methodology} \label{n4bures} To continue the full line of research reported here for the cases $n=2, 3$, to $n > 3$, it would be useful to extend the work of Byrd and Sudarshan \cite{byrd,byrdsudarshan} on the Euler angle parameterization of $SU(3)$ to such higher $n$. However, computation of the Hall constants $C_{n}$ (\ref{hallconstant}) does not depend on parameterizations of $SU(n)$. We have, in fact, been able to obtain exceedingly strong {\it numerical} evidence that $C_{4}$ is, in fact, equal to ${71680 / \pi^{2}} \approx 7262.7$. Let us, first, mention some methodological considerations useful in deriving this result (and, in general, $C_{n}$, $n > 4$). The parameterizations of Boya {\it et al} \cite{boya} of the diagonal $2 \times 2$ and $3 \times 3$ matrices differ, in that in the $2 \times 2$ case (\ref{two}) only matrices in which the (1,1)-entry is at least as great as the (2,2)-entry are generated (due to the restriction of the angular parameter, $\theta$, to the range $[0, \pi /2]$), while in the $3 \times 3$ case (\ref{three}), no order is imposed on the diagonal entries (the parameters $\theta$ and $\phi$ both varying freely between 0 and $\pi$). Now, in performing (the apparently necessary) numerical (as opposed to symbolic) integrations to obtain the Hall constants $C_{n}$ $(n > 3)$, it seems to be considerably more computationally effective to integrate over only those diagonal matrices in which (say) the (1,1)-entry is no less than the (2,2)-entry, which in turn is no less than the (3,3)-entry, {\it etc.} (This helps to minimize troublesome oscillations.) Then, the result can be multiplied by the number ($n!$) of permutations of $n$ objects to yield $C_{n}$, since the result of the integration must be invariant under any other of the $n!-1$ possible orderings (permutations) that can be imposed on the diagonal entries of the $n \times n$ diagonal matrices. In precisely this manner, we were able to obtain (using the numerical integration of interpolating function command of MATHEMATICA) the result $71680.000001/ \pi^{2}$ for $C_{4}$. This we take as overwhelming evidence that $C_{4}$, in fact, equals $71680 / \pi^{2}$, particularly so, since 71,680 has the highly structured prime decomposition of $2^{11} \cdot 5 \cdot 7$. (It is also interesting to note that $35 = 5 \cdot 7$ appears in the numerator of $C_{3}$, that is, $35 / \pi$, or, to the same effect, $C_{4} = 2^{11} C_{3}/ \pi$. To illustrate the procedure followed, let us first parameterize the $4 \times 4$ nonnegative diagonal matrices of trace unity in the following fashion (cf. (\ref{two}), (\ref{three})), \begin{equation} D = \pmatrix{ \cos^{2}{\theta / 2} & 0 & 0 & 0 \cr 0 & \sin^{2}{\theta / 2} \cos^{2}{\phi / 2} & 0 & 0 \cr 0 & 0 & \sin^{2}{\theta / 2} \sin^{2}{\phi / 2} \cos^{2}{\zeta / 2}& 0 \cr 0 & 0 & 0 & \sin^{2}{\theta / 2} \sin^{2}{\phi / 2} \sin^{2}{\zeta / 2} \cr}. \end{equation} Then, the (truncated) region of integration employed above --- corresponding to successively nonincreasing diagonal entries --- can be described as $\zeta \in [0, \pi / 2], \phi \in [0, f(\zeta)], \theta \in [0, f(\phi)]$, where $f(x) = 2 \cot^{-1}{(\cos{x/2})}$, rather than $\zeta, \phi, \theta \in [0, \pi]$, as in the apparently suggested parameterization of Boya {\it et al} \cite{boya}, which would yield all possible diagonal matrices, without regard to the ordering of their elements. If we construct a similar truncated region of integration in the case $n=3$, then we find that the expected values of the eigenvalues are .802393, .181878 and .0157299. In the $n=2$ case, the analogous values are (as previously noted), $1/2 \pm 4 / 3 \pi$, that is .924413, and .0755868. \subsection{The Hall constant for $n=5$} We have also attempted to compute $C_{5}$, in the manner of sec.~\ref{n4bures}, with the use of MATHEMATICA. We obtained (using the Gauss-Kronrod integration method with a working precision of twenty-one digits, rather the machine precision of sixteen) the result $2342475135.00/ \pi^{2}$. Now, it is most interesting to note (particularly, in light of our results for $C_{n}$, $n<5$, that \begin{equation} \label{C5} 2,342,475,135 = 21 \Pi_{i=2}^{9} P_{i} = 21 (3 \cdot 5 \cdot 7 \cdot 11 \cdot 13 \cdot 17 \cdot 19 \cdot 23), \end{equation} where $P_{i}$ denotes the $i$-th prime number (taking the sequence of primes to be $2, 3, 5, \ldots$). Thus, we have acquired strong evidence that, in fact, $C_{5} = 2342475135/ \pi^{2}$. (As a simple exercise, we looked at the one hundred thousand consecutive integers containing 2,342,475,135 as their midpoint, and computed all their prime decompositions. All the others had at least one prime factor greater than 23.) The seemingly independent factor of 21 in (\ref{C5}) will also apparently be found below in the (odd) case $n=7$. \subsubsection{Prime factorials} In \cite{caldwell}, the product of the primes less than or equal to $p$ is denoted $p \#$. (The issue there, as in several of the works cited there, was to test $p \# \pm 1$ for primality (cf. \cite{shor}).) Let us point out a 1952 article \cite{duarte}, entitled ``Tables of logarithms of the prime factorials from 2 to 10007'' (a synopsis of which can be found in Mathematical Reviews 16, 112f). It is noted that by the prime number theorem \cite{primenumbertheorem} the ratio of the sum of the logarithms of the primes from 2 to $p$ to $p$ itself approaches 1 as $p \rightarrow \infty$. So, if it eventuates that the general formula for $C_{n}$, at least for odd $n$, contains a term of the form $p(n) \#$, where $p(n)$ is a prime as a function of $n$, which grows indefinitely large with $n$ itself, then it should be possible to asymptotically replace $p(n) \#$ by $e^{p(n)}$. (``A version of the prime number theorem states that the product of the primes less than $x$ is asymptotically $e^{x}$ [citing the well-known treatise \cite[Theorem 434]{hardy}], but the error term is notoriously large, so it is probably unrealistic to expect to be able to compute far enough to get within the necessary epsilon'' \cite{guy}.) We also observe that with the use of Wilson's theorem \cite{wilson}, $(p-1) ! \equiv -1 \mbox{mod} {p}$, one could express $p \#$ in terms of the (more) standard factorial function $p!$. (In \cite{mat}, the primes are {\it defined} in terms of factorials.) Let us point out that Ellinas and Floratos have recently studied the ``prime decomposition'' of an $n \times n$ density matrix into a sum of separable density matrices with dimensions determined by the coprime factors of $n$ \cite{ellinas} (cf. \cite{number}). \subsection{Preliminary investigations of the Hall constant for $n>5$, with the use of quasi-Monte Carlo integration} \subsubsection{$n=6$} \label{subn6} Of course, as the dimensionality of the $n$-state quantum systems increases (that is, $n$ itself increases), the numerical integrations required to sufficiently narrow estimates of the corresponding Hall constant become increasingly more difficult, and it is hard to judge what is precisely the optimum numerical/programming strategy to employ. Following the methodology outlined in sec.~\ref{n4bures}, based on the ordering of the eigenvalues, MATHEMATICA did yield (using the standard default options) an estimate of $C_{6}$, representable in the form, $1.4616286 \cdot 10^{16} / \pi^{3}$ (although diagnostics as to inadequate precision were issued during the course of the computation). This result, coupled with our observation of the pattern of $C_{n}$ for $n < 6$, might lead us to speculate that the numerator of $C_{6}$ is either the seventeen-digit number, $14,616,907,579,654,144 = 2^{41} \cdot 17^{2} \cdot 23$ or $14,623,504,649,420,800 = 2^{42} \cdot 5^{2} \cdot 7 \cdot 19$, with the denominators, in both cases, being $\pi^{3}$. When we employed the quasi-Monte Carlo (Halton-Hammersley-Wozniakowski) procedure \cite[chap. 3]{hammersley} \cite{saar} of MATHEMATICA, to numerically integrate over a hypercube of volume $\pi^{5}$ (corresponding, thus, now to no particular distinguished ordering of the six eigenvalues), we obtained (with {\it no} diagnostics at all being generated in two separate analyses --- having set maxima of ten and fifty million sample points) a result of the form $1.536355674 \cdot 10^{16} / \pi^{3}$. To a very high accuracy, this numerator can be approximated by $15,363,556,773,986,304 = 2^{28} \cdot 3^{4} \cdot 13^{2} \cdot 37 \cdot 113$ for the numerator of the presumptive value of $C_{6}$ (with denominator, again, $\pi^{3}$). Sloan and Wo\'zniakowski have noted that recently ``Quasi-Monte Carlo algorithms have been successfully used for multivariate integration of high dimension $d$, and were significantly more efficient than Monte Carlo algorithms'' \cite{sloan}. However, in comparing the two sets of results here for $n=6$ it is very important to bear in mind that MATHEMATICA sets its precision and accuracy objectives much lower when Monte Carlo procedures are employed \cite{saar}. (Of course, these default values can be reset, but in the preliminary analyses reported here, they have not been, though we intend to do so in future studies.) \subsubsection{$n=7$} \label{subsecn7} The quasi-Monte Carlo MATHEMATICA procedure produced an estimate (with no accompanying diagnostics, having set a maximum of four million sample points) of $C_{7}$ as $2.4811899 \cdot 10^{24} / \pi^{3}$. The numerator of this fraction (cf. (\ref{C5})) could be approximated (to a relative error of less than three-tenths of one percent) by the product of 21 and the (eighteen) consecutive primes from 2 to 61 (in the notation of \cite{caldwell}, this is $61 \#$), that is, $C_{7} = 21 \Pi_{i=1}^{18} P_{i} / \pi^{3}$. However, this degree of accuracy does not at all seem satisfactory, particularly in light of the proximate results for $n=6$ and 8, though odd $n$'s appear to present greater computational challenges. \subsubsection{$n=8$} The quasi-Monte Carlo procedure estimated $C_{8}$ as $4.1836777028 \cdot 10^{35} / \pi^{4}$ (though, unlike the $n=6,7$ cases, a failure to converge was reported --- that is, with the preassigned use of at most ten million sample points). Nonetheless, this outcome can be fit to a very high accuracy (less than one part in one hundred million) by taking $C_{8}$ to be $2^{89} \cdot 3^{2} \cdot 5 \cdot 13^{2} \cdot 31 \cdot 47 \cdot 61 / \pi^{4}$. \subsection{$n \geq 9$} For the case $n=9$, convergence was not obtained with the use of six million sample points. The result given was $7.631832917 \cdot 10^{47} / \pi^{4}$. We have also made tentative attempts, using the quasi-Monte Carlo procedure again, to estimate $C_{n}$ for $n =10$ and 11 (but it appears that considerable investment of computer resources is needed for sufficiently satisfactory answers). For $n=10$, based on a maximum of two million sample points, convergence was not obtained and the result $6.334733996 \cdot 10^{62} / \pi^{5}$ reported. (No related decomposition was immediately apparent.) For $n=11$ and 12, using a maximum of one million sample points in both cases, the results $3.254198489 \cdot 10^{78} / \pi^{5}$ and $5.218327334 \cdot 10^{97}/ \pi^{6}$ were gotten (without convergence or any obvious associated simple prime decompositions, however). Our impression is that the computations are considerably more difficult for the odd values of $n$ than for the even, which may be some reflection of the simpler formulas displayed above for even $n$ (cf. \cite{chen,cvij}). (For $n>7$, the MATHEMATICA compiler was unable to handle the small numbers appearing in the calculation, and then proceeded with the use of the uncompiled evaluation, leading to slower running times \cite{saar}.) Of course, there exists a wide range of possible approaches to numerical integration problems of this kind, including, certainly, the use of alternative programming languages, in particular, FORTRAN. The trade-offs between these various options need to be assessed. As a reference point, against which one can attempt to compare the (reciprocals of the) several values of $C_{n}$ above, let us recall that the area of an $n$-sphere of radius $r$ is given by $2 \pi^{n/2} r^{n-1} / \Gamma(n/2)$ \cite{fusaro}. Also of similar interest is Euler's formula \begin{equation} \label{riemann} \zeta(2 n) = {(-1)^{n-1} \over 2 (2 n)!} (2 \pi)^{2 n} B_{2 n}, \quad n =1,2,3,\ldots \end{equation} where $\zeta(s)$ is the Riemann zeta function, $\sum_{n \geq 1} n^{-s}$, and $B_{n}$ is the (necessarily rational) $n$-th Bernoulli number \cite[Vol. I, pp. 75 and 211]{terras}. (The values of $\zeta(n)$ for positive odd integers $n$, however, have not been expressed in such a simple form \cite[Vol. III, p. 1695]{ito}. Infinite series representations are, in fact, reported in \cite{chen,cvij}.) Initial attempts to find an explanatory formula for the sequence of (integral) numerators of $C_{n}$, using on-line programs of C. Krattenthaler (``Rate'') \cite[App. A]{advanced} and of N. Sloane (``superseeker'') \cite{sloane}, did not succeed. However, we did eventually find a somewhat intriguing connection (further buttressed by some related analyses, discussed below in sec.~\ref{further}) between the results here and sequence A035077 of \cite{sloane}, which gives the denominators of partial sums of $B_{2 n}$. The numerator of: (1) $C_{2}$ is (rather trivially) twice the first entry in this sequence; (2) $C_{3}$ is one-half the fourth entry; (3) $C_{4}$ is $2^{10} = 1024$ times the fourth entry; and (4) $C_{5}$ is 63 times the thirteenth entry of A035077. (We note that Donaldson has found simple proofs of various formulas for {\it symplectic} volumes involving Bernoulli numbers \cite{donaldson,witten,jeffrey}.) \subsection{Average von Neumann entropy of $n$-state quantum systems with respect to Bures prior probability distributions} \label{ENTROPY} As one application of these computations of the Hall constants ($C_{n}$), let us note that with respect to the Bures probability distribution (\ref{n2}) the average von Neumann entropy, $-\mbox{Tr} \rho \log{\rho}$, is exactly $2 \log{2} - 7/6 \approx .219628$ nats for the two-state systems and, now using numerical integration, .507937 nats for the three-state systems (cf. \cite{page}). (Since we employ the natural logarithm here, the unit of information is the {\it nat}, which is equivalent to $1/ \log{2} \approx 1.4427$ {\it bits}.) This latter result, to a high degree of precision --- that is, to ten significant places, can, in fact, be written as $3 \log{3} - 3917/1405$. We also computed the Bures average entropy for the four-state quantum systems, obtaining .751771 nats. To eight significant places, this can be written as $4 \log{4} - 32135/6704$. Similarly, for $n=5$, using our knowledge of $C_{5}$, we obtain an average entropy of .954103 nats. This is closely approximated (to at least nine places, according to our calculations) by $5 \log{5} - 40045/5648$. \subsection{Auxiliary analyses of variations of Hall integrals and the role of Bernouilli numbers} \label{further} Since it appeared to be quite challenging to determine the Hall constants ($C_{n}$) for $n > 5$, we thought that it might be revealing (possibly helpful in deriving a general formula for arbitrary $n$), as well as being of independent interest, to investigate more tractable variations. To do so, we replaced the exponent two in (\ref{hallformula}) and (\ref{hallconstant}) by either one (corresponding to {\it real} quantum systems) or four (for quaternionic quantum systems) \cite{mehta}, and other positive integer values of less immediate {\it physical} interest, as well. For the $n=3$ (spin-1) quaternionic case (that is, using an exponent of four), the counterpart of the previously derived (sec.~\ref{proximate}) Hall constant $35 / \pi $ is $1616615/ {226 \pi} = 5 \cdot 7 \cdot 11 \cdot 13 \cdot 17 \cdot 19 / 226 \pi$. (The numerator here is one-half the tenth member of the sequence A035077, comprised of denominators of partial sums of the Bernoulli numbers $B_{2 n}$ \cite{sloane}.) When we use an exponent of six, the result is $ 100280245065 /88252 \pi = 3 \cdot 5 \cdot 7 \cdot 11 \cdot 13 \cdot 17 \cdot 19 \cdot 23 \cdot 29 \cdot 31 / 88252 \pi$. (This numerator is precisely the seventeenth member of A035077.) If we use an exponent of one (corresponding to the case of real quantum systems), the result is $1 /4 \pi$, and if we employ an exponent of three, the normalization constant is $ 105 / 128 \pi = 3 \cdot 5 \cdot 7 / 128 \pi$. (The numerator here is the fifth member of A035077.) For an exponent of five, we have $15015 / 8192 \pi = 3 \cdot 5 \cdot 7 \cdot 11 \cdot 13 / 8192 \pi$, the numerator being three times the seventh entry of Sloane's sequence. For an exponent of seven, the result was the eleventh member of the sequence A035077, that is, 969969, divided by $262144 \pi$. When an inquiry was made of Neil Sloane as to whether to his knowledge there were any published discussions of this sequence, he replied ``No, I was just looking at various sequences of important rationals, and thought that the pair A035078/A035077 should be in the database''. (However, he later pointed out that the von Staudt-Clausen Theorem \cite[p. 10]{rademacher} was relevant to questions involving sums of Bernoulli numbers.) For the case $n =3$, MATHEMATICA rejected our efforts to compute any further exact integrals having integer exponents greater than seven. (It would appear that the use of an exponent of eight would be associated with the {\it octonionic} quantum systems \cite{deleo}.) For the analogous set of variations with $n=2$, use of odd exponents in (\ref{hallformula}) and (\ref{hallconstant}) lead to divergent results. For an exponent of two, the result is $2 \pi$, for four, $8 / 3 \pi $, for six, $16 / 5 \pi $, for eight, $128 / 35 \pi $, \ldots. For comparable scenarios based on $n=4$, we were unable to proceed with exact integrations. Our numerical computation of the analog of the Hall constant employing an exponent of unity (the real quantum case), yielded $ \pi^{2} / .079271$. But we were unable to determine if this result bore any relation to the sequence A035077. \section{{\it QUASI}-BURES PROBABILITY DISTRIBUTIONS OVER THE $n \times n$ DENSITY MATRICES} \label{QUASI} In line with the work reported in \cite{kratt}, it would be of interest to obtain formulas for the averages over the eight-dimensional convex set of $3 \times 3$ density matrices ($\rho$) with respect to the Bures prior probability distribution (\ref{new}) of the $m$-fold tensor products of $\rho$. As $m \rightarrow \infty$, the relative entropy of these products with respect to the averaged $3^{m} \times 3^{m}$ density matrix gives us the (Bures) asymptotic redundancy for the universal quantum coding of {\it three}-state systems. \subsection{The quasi-Bures case $n=2$} In \cite{kratt} and further yet unreported work, the Bures prior probability density (\ref{n2}) for the two-state systems was found to closely resemble the (what we term ``quasi-Bures'') probability density, \begin{equation} \label{wsquiggle} \mbox{d} p_{quasi:2}(\theta,\alpha,\beta) = .226231 (\tan^{\sec{\theta}}{\theta / 2}) \cos{\theta} \cot{\theta} \sin{\beta} \mbox{d} \theta \mbox{d} \alpha \mbox{d} \beta, \qquad 0 \leq \theta \leq {\pi \over 2}, \quad 0 \leq \alpha \leq 2 \pi, \quad 0 \leq \beta \leq \pi \end{equation} which yields, it appears, {\it both} the asymptotic minimax and maximin redundancies (as ``Jeffreys' prior'' \cite{kass} does classically \cite{cb1,cb2}). This common value, if one ignores the error term, as appears to be legitimate, is $(3 \log{m}) / 2 - 1.77062$, while the Bures probability distribution (\ref{n2}) has been shown to be, incorporating the error term, associated with an asymptotic redundancy of $(3 \log{m}) /2 - 1.77421 + O(1/\sqrt{m})$ \cite[p. 29]{kratt}. In general, for any probability distribution $w(\theta), \theta \in [0, \pi/2]$, the asymptotic redundancy for the two-state quantum systems takes the form, \begin{equation} \label{redundancy} {3 \over 2} \log{{m \over 2 \pi}} -{1 \over 2} -2 \log{\sin{\theta}} + 2 (\sec{\theta}) \log{\tan{{\theta \over 2}}} -\log{w(\theta)} + o(1). \end{equation} Standard variational arguments can, then, be used to show (ignoring the error term, the legitimacy of which seems plausible, but has not yet been rigorously justified) that the particular $w(\theta)$ yielding both the maximin and minimax redundancies is simply proportional to the $\theta$-dependent part of $p_{quasi:2}(\theta,\alpha,\beta)$, indicated in (\ref{wsquiggle}). The reciprocal of the corresponding ``Morozova-Chentsov function'' $c(x,y)$ \cite{petzsudar,petzlaa} for (\ref{wsquiggle}) is the exponential or identric mean \cite[eq. (1.3)]{qi} of $x$ and $y$, \begin{equation} \label{qi13} I(x,y) = e^{-1} (x^{x}/y^{y})^{{1 \over x-y}},\quad x \ne y \end{equation} ($I(x,x)= x$), while for the Bures (minimal monotone) metric, it is the (more commonly encountered) arithmetic mean $(x+y)/2$. The associated operator monotone functions \cite{petzsudar,petzlaa} are $f(t) =(1+t)/2$ for the Bures metric, and $f(t) = t^{t/(t-1)}/e$, for the metric giving (\ref{wsquiggle}). (The Morozova-Chentsov functions fulfill the relation $c(x,y) =1/yf(x/y)$.) Perhaps the {\it exponential} mean arises in this context because the von Neumann entropy is the {\it logarithmic} relative entropy, and of course the exponential and logarithmic functions are inverses of one another. This leads us to speculate that if one were to employ, following \cite{lesniewski}, the ``quadratic relative entropy'' or the ``Bures relative entropy'' instead, then, in the parallel universal coding context, the minimax/maximin would be achieved by the means corresponding to the new forms of inverse functions. While the logarithmic relative entropy is based on the operator convex function, $g(t)=-\log{t}$, the quadratic form relies upon $g(t) = (t-1)^{2}$ and the Bures form on $g(t) = (t-1)^2/(t+1)$ \cite{lesniewski}. Here, $f(t) = (t-1)^{2}/(g(t) + t g(1/t))$. In Fig.~\ref{comparison}, we jointly display the univariate marginal probability distributions of (\ref{n2}) and (\ref{wsquiggle}), revealing that they closely resemble one another, with the quasi-Bures distribution assigning relatively greater probability to the states more pure in character ($\theta \leq .443978$). \begin{figure} \centerline{\psfig{figure=comparison.eps}} \caption{One-dimensional marginals of the ($n=2$) Bures probability distribution (\ref{n2}) and the quasi-Bures distribution (\ref{wsquiggle}) (assuming greater values for $\theta \in [0,.443978]$), which yields both the asymptotic minimax and maximin redundancies for the universal quantum coding of two-state systems} \label{comparison} \end{figure} \subsection{The quasi-Bures case $n=3$} The three-state counterpart of (\ref{wsquiggle}) (that is, the probability distribution associated with the exponential/identric mean, rather than the arithmetic mean, as for such Bures distributions) is \begin{equation} \label{quasiBures} \mbox{d} p_{quasi:3}(\theta,\phi,\alpha,\beta,\gamma,\kappa,a,b) = .000063495 u ({\tan^{(1 + \sec{\phi})}}{\phi / 2}) {\csc^{4}{{\theta \over 2}}} {\csc^{6}{{\phi \over 2}}} \end{equation} \begin{displaymath} (\cos{{\phi \over 2}} \tan{{\theta \over 2}})^{{16 {\cos^{2}{{\phi \over 2}}} {\sin^{2}{{\theta \over 2}}} \over {v+w}}} (\sin{{\phi \over 2}} \tan{{\theta \over 2}})^{2 -{8 (1 + \cos{\theta}) \over {w -v}}} \mbox{d} \theta \mbox {d} \phi \mbox{d} \widetilde{\Omega}_{3}, \end{displaymath} where $u$ is given in (\ref{newvariable}), $\mbox{d} \widetilde{\Omega}_{3}$ in (\ref{conditionalhaar}) and \begin{equation} v = 2 + 6 \cos{\theta}, \quad w = \cos{(\theta-\phi)} -2 \cos{\phi} +\cos{(\theta +\phi)}. \end{equation} The corresponding two-dimensional marginal probability distribution over the variables $\theta$ and $\phi$ is exhibited in Fig.~\ref{exponentialspin1}. \begin{figure} \centerline{\psfig{figure=exponentialprior2.eps}} \caption{Bivariate marginal of the quasi-Bures probability distribution (\ref{quasiBures}) over the variables $\theta$ and $\phi$, for three-state systems} \label{exponentialspin1} \end{figure} As would be anticipated from Fig.~\ref{comparison}, this figure closely resembles Fig.~\ref{twodim}. In Fig.~\ref{plotdifference}, we show the result obtained by subtracting the bivariate marginal Bures probability distribution shown in Fig.~\ref{twodim} from its quasi-Bures counterpart in Fig.~\ref{exponentialspin1}. \begin{figure} \centerline{\psfig{figure=difference.eps}} \caption{Difference obtained by subtracting the three-state Bures bivariate marginal probability density in Fig.~\ref{twodim} from the quasi-Bures one, associated with the exponential/identric mean, displayed in Fig.~\ref{exponentialspin1}} \label{plotdifference} \end{figure} \subsection{Quasi-Bures counterparts of Hall (Bures) normalization constants} If we replace the term $d_{j} + d_{k}$, occurring in the denominator of the expression (\ref{hallconstant}), by (twice) the exponential/identric mean (\ref{qi13}) of $d_{j}$ and $d_{k}$, that is, $I(d_{i},d_{j})$, the resultant expression becomes a formula for $\tilde{p}_{quasi:n}(d_{1},\ldots,d_{n})$, now interpreting $C_{n}$ to be the normalization constant for the corresponding quasi-Bures probability distribution. For $n=2$ then (taking $\theta \in [0,\pi]$ and not $[0,\pi/2]$ as in \cite{boya}), rather than $2 / \pi \approx .63662$, we obtain .769427, and for $n=3$, instead of $35 / \pi \approx .089759$, we find .138681. \section{CONCLUDING REMARKS} We hypothesize that the (full eight-dimensional) quasi-Bures probability distribution (\ref{quasiBures}) associated with Fig.~\ref{exponentialspin1} will furnish the common asymptotic (minimax and maximin) redundancies for universal quantum coding in that higher-dimensional setting, paralleling the result (not yet fully formally demonstrated, however) for the three-dimensional convex set of two-state systems (cf. \cite{jozsa}). In this regard, it might prove computationally convenient, as a heuristic device, to replace the quasi-Bures probablity distributions by their (closely approximating) Bures analogues, since certain exact (symbolic) integrations are achievable (at least, for the cases $n = 2, 3$) with the Bures distributions, but apparently not with the quasi-Bures ones, for which numerical methods seem to be necessary. Although a parameterization of SU(4) is not relevant, as already noted, to the computation of the Hall constant $C_{4}$, that is, $71680 / \pi^{2}$, and to that of the corresponding average (von Neumann) entropy (sec.~\ref{ENTROPY}), it would be essential in investigating the universal coding of four-state quantum systems, since the tensor products of $4 \times 4$ density matrices have to be calculated and it would, therefore, be necessary to implement formula (\ref{one}). In a personal communication, M. Byrd has indicated that he has undertaken the (challenging) task of developing such an (Euler angle) parameterization of $SU(4)$. Such a parameterization of $SU(4)$ --- in conjunction with the knowledge, acquired here, of $C_{4}$ --- might also prove of value in estimating the volume of $2 \times 2$ separable quantum states \cite{zhsl,zyczrecent} and lead to numerically more stable results than those reported in \cite[Table I]{slaterjpa2}, since the ``over-parameterizations'' of the unitary matrices used there could then be avoided, due to the ``dropping out'' (as pointed out for $n=2,3$ immediately after (\ref{ranges}) above) of certain Euler angles in the formation of the product (\ref{one}). We also note that in \cite{slaterjpa2} (cf. \cite{slaterduan}), the Bures (minimal monotone) metric was found to yield higher {\it a priori} probabilities of entanglement than other monotone metrics (in particular, the Kubo-Mori-Bogoliubov and maximal ones). Presumably, even in any computationally improved form of analysis, this conclusion would be unaltered. We have also pursued a traditional (pseudo-random number) Monte Carlo approach to estimating the Hall normalization constants for $2 \leq n \leq 16$. However, the degrees of precision attained were not satisfactory. (For discussions of the comparative computational complexities of the pseudo- and quasi-Monte Carlo methods, see \cite{sloan,frank}.) We are presently attempting to obtain a more precise estimate of the Hall constant $C_{6}$, in particular, using {\it non}-Monte Carlo (that is, adaptive) integration methods. Our best current estimate of $C_{6}$ is $1.534836628 \cdot 10^{16} / \pi^{3}$. The numerator can be very well approximated by $15348366279966720 = 2^{33} \cdot 3 \cdot 5 \cdot 7^{2} \cdot 11 \cdot 13 \cdot 17$, which seems more satisfactory than the results reported in sec.~\ref{subn6}, based on the quasi-Monte Carlo procedure. Now, it is most interesting to note that this numerator is precisely $2^{33} \cdot 7$ times the ninth entry of the sequence A035077, we have repeatedly referenced above. We are also compelled to observe that our educated conjecture as to the numerator of $C_{7}$ (sec.~\ref{subsecn7}) is exactly forty-two times the thirty-third entry --- which we had to compute ourselves, since Sloane's published list does not extend this far --- of A035077. \acknowledgments I would like to express appreciation to the Institute for Theoretical Physics for computational support in this research and to M. Byrd and K. \.Zyczkowski each for a number of helpful communications, as well as to C. Krattenthaler for his insightful analyses, and to M. J. W. Hall for pointing out to me an erroneous statement in an earlier version. Also I thank J. Stopple for the reference to \cite{terras} and his interest in this work and to M. Choptuik for a discussion concerning the relative merits of various numerical integration routines.
\section{ Introduction} The birefringence of $\gamma$-quanta with energies $ >1 $ GeV propagating in single crystals was predicted in \cite{C}. The main process by which $\gamma$-quanta are absorbed in single crystals is the electron-positron pair production. The cross section of the process depends on the direction of linear polarization of the $\gamma$-quanta relative to the crystallographic planes. As a result of interaction with the electric field of the single crystal, a monochromatic, linearly polarized beam of $\gamma$-quanta comprises two electromagnetic waves with different refractive indices, so that linear polarization is transformed into circular polarization or vice versa. This polarization phenomenon would be observed for symmetric orientations of single crystals with respect to the direction of motion of $\gamma$-quanta. The general case of the propagation of $\gamma$-quanta in single crystals was considered in \cite{MMF,MV1,MV2}. In these papers it was shown that the propagating $\gamma$-beam is a superposition of the two elliptically polarized waves and unpolarized $\gamma$-beam obtain some degree of circular and linear polarization after passage through a single crystal. In case, describing in \cite{C}, the beam of $\gamma$-quanta is a superposition of the two linearly polarized waves and unpolarized beam obtain only some degree of linear polarization after propagation in single crystals. It is important to note that no experiments have been performed to date to corroborate the transformation of $\gamma$-beam polarization in single crystals, despite the notable lapse of time since the publication of \cite{C}. It is at least two essential purposes for experimental investigations of the birefringence in single crystals. There are: 1)The nature of phenomenon is a manifestation of the nonlinearity of Maxwell's equations for the electromagnetic vacuum. Of course, a single crystal contain carriers of electric charge (electron, ions, etc), but their direct presence is significant only if the frequencies of the electromagnetic radiation passing through the single crystal are low, while at high frequencies the fields formed by these charges play the main role. Thus the observation of the birefringence in single crystals is indirect experimental proof of existence of the similar effect in electromagnetic vacuum (see \cite{WDHG} and the literature cited therein); 2) Some possibilities exist to utilize this phenomenon in experiments on modern accelerators (see \cite{MV2,BT,KP} and the literature cited therein). \section{ Refractive indices of $\gamma$-quanta in single crystals.} Now we have found the refractive indices of $\gamma$-quanta propagating in a single crystal. Below we rewrite the components of a complex permittivity tensor $\varepsilon_{ij}= \varepsilon'_{ij}+\varepsilon''_{ij} , i, \, j =1,2$ (the process is determined by the transverse part of the tensor) from paper \cite{MMF}. Let us consider a high-energy beam of $\gamma$-quanta moving at a small angle $\theta$ with a reciprocal lattice axis defined by vector $\bf{G_1}$. Then in the Cartesian system of coordinates such that one axis is oriented approximately parallel to the direction of motion of the $\gamma$-quanta and other two axes lie in planes determined by the vectors $\bf{G_1},\, \bf{G_2}$ and $\bf{G_1}, \, \bf{G_3}$, the tensor $\varepsilon_{ij}$ is a sum over reciprocal lattice vectors ${\bf{g}}= n_2 {\bf{G_2}}+n_3 {\bf{G_3}}\, (n_1=0, \theta \ll 1)$ and has the following components: \begin{eqnarray} \varepsilon'_{11}={S^{'}\over 2}+ {{BN\bar{\sigma}}\over {8\pi}} {\hbar \over mc} \sum_{\bf{g}} \Phi(g)\,(g_2^2-g_3^2)z_g^2F'_1(z_g). \nonumber \\ \varepsilon'_{22}={S^{'}\over 2}- {{BN\bar{\sigma}}\over {8\pi}} {\hbar \over mc} \sum_{\bf{g}} \Phi(g)\,(g_2^2-g_3^2)z_g^2F'_1(z_g). \nonumber \\ \varepsilon'_{12}=\varepsilon'_{21}= + {{BN\bar{\sigma}}\over {8\pi}} {\hbar \over mc} \sum_{\bf{g}} \Phi(g)\,(2g_2g_3)z_g^2F'_1(z_g). \label{1} \\ S'= 2+{{BN\bar{\sigma}}\over {\pi}}{\hbar \over mc} \sum_{\bf{g}} \Phi(g)\,(g_2^2+g_3^2)z_g^2F'_2(z_g,1). \nonumber \\ z_g= {{2mc^2} \over {E_{\gamma}\theta(g_2\cos\alpha +g_3 sin\alpha)}} = {1 \over { n_2W_V +n_3W_H}}. \label{2} \end{eqnarray} The summation over {\bf{g}} satisfies the condition \begin{equation} z_g>0. \label{3} \end{equation} \begin{eqnarray} \varepsilon''_{11}= {S''\over 2} - {BN \bar{\sigma} \over 16} {\hbar \over mc} \sum_{\bf{g}} \Phi(g)(g_2^2 - g_3^2)F''_1(z_g), \nonumber \\ \varepsilon''_{22}= {S''\over 2} + {BN \bar{\sigma} \over 16} {\hbar \over mc} \sum_{\bf{g}} \Phi(g)(g_2^2 - g_3^2)F''_1(z_g), \nonumber \\ \varepsilon''_{12}=\varepsilon''_{21}= - {BN \bar{\sigma} \over 16} {\hbar \over mc} \sum_{\bf{g}} \Phi(g)(2g_2 g_3)F''_1(z_g), \label{4} \\ \nonumber S''= \varepsilon_A+{{BN\bar{\sigma}}\over {2}}{\hbar \over mc} \sum_{\bf{g}} \Phi(g)\,(g_2^2+g_3^2)z_g^2F''_2(z_g,1). \end{eqnarray} The summation over {\bf{g}} satisfies the condition \begin{equation} 0<z_g \le 1 \label{5} \end{equation} The functions $F'_1, F'_2, F''_1, F''_2, $\footnote{ Note that these functions are also used for description of the birefringence in the laser electromagnetic wave \cite{MV1,MV}} are equal to: \begin{equation} {F_1}^{\prime }(z) = \cases{ [\sqrt{1-z}+{z\over 2}L_-]^2+[\sqrt{1+z}-{z\over 2}L_+]^2 - {{\pi^2z^2}\over {4}},\; 0<z \le 1, \cr -[\sqrt{z-1} -z \arcctg\sqrt{z-1}]^2+[\sqrt{1+z}- {z\over 2}L_+]^2 , \; z>1. \cr} \label{6} \end{equation} \begin{equation} {F_2}^{\prime }(z,\mu ) = \cases{ -2-\mu -(1+\mu (z-{z^2\over 2})){1\over 4} L^2_- -(1-\mu (z+{z^2\over 2})){1\over 4}L^2_+ + \cr +{{(1+\mu z )\sqrt{1-z}}\over{2}}L_- -{{(\mu z-1)\sqrt{z+1}}\over {2}}L_+ + {\pi^2 \over 4}(1+\mu (z-{{z^2}\over 2})),\; 0<z \le 1, \cr -2-\mu +(1+\mu (z-{z^2\over 2})) \arcctg^2(\sqrt{z-1}) -(1-\mu (z+{z^2 \over 2})){1\over 4} L^2_+ + \cr + (1+\mu z)\sqrt{z-1} \arcctg\sqrt{z-1} -{{(\mu z-1)\sqrt{1+z}}\over{2}} L_+ ,\; z>1. \cr } \label{7} \end{equation} \begin{equation} {F_1}^{\prime \prime}(z)= \cases{ z^4 (L_- + {{2\sqrt{1-z}}\over{ z}}) , \; 0< z \le 1, \cr 0, \; z>1. \cr} \label{8} \end{equation} \begin{equation} {F_2}^{\prime \prime}(z,\mu )= \cases{ z^2((1+\mu (z-{z^2\over 2}))L_- -\sqrt{1-z}(1+\mu z)), \; 0<z \le 1, \cr 0, \; z> 1 \cr} \label{9} \end{equation} The functions $L_+,\, L_- $ are equal to: \begin{equation} L_+ =ln{{\sqrt{1+z}+1}\over {\sqrt{1+z}-1}} \, . \nonumber \label{10} \end{equation} \begin{equation} L_- =ln{{1+\sqrt{1-z}}\over {1-\sqrt{1-z}}} \, . \label{11} \end{equation} In these equations $E_\gamma$ is the energy of $\gamma$-quanta, $m$ is the electron mass, c is the speed of light, $\alpha$ is the angle between planes $({\bf{G_1}}, {\bf{G_2)}}$ and $({\bf{G}_1}, {\bf{K}})$, where ${\bf{K}}$ is the momentum of $\gamma$-quanta. The value $\Phi(g)$ is determined by following relation: \begin{eqnarray} \Phi(g)= |S({\bf{g})}|^2(1-F(g))^2\exp^{-Ag^2}/g^4 , \label{12} \end{eqnarray} where S(g) is the structure factor, F(g) is the form factor of an atom in the single crystal and A is the mean-square amplitude of thermal vibrations of the atoms. N is the number of atoms per unit of volume. \begin{eqnarray} B= {{16\pi^2} \over {N_S \Delta}}, \qquad \qquad \bar{\sigma}=\alpha_e Z^2 r_e^2, \label{13} \end{eqnarray} where $ \alpha_e$ is the fine-structure constant, $r_e$ is the classical electron radius, $Z$ is the atomic number of the material of the single crystal, $\Delta$ is the volume of the elementary cell and $N_S$ is the number of atoms per this cell. The term $\varepsilon_A$ in Eqs.(5) takes into account the absorption of $\gamma$-quanta on the thermal vibrations of the lattice and is equal to \begin{equation} \varepsilon_A= {\bar{\sigma} Nc \hbar \over E_\gamma} ({2 \over 3} \psi^{am}_1 + {1 \over 9} \psi^{am}_2), \label{14} \end{equation} where the values $\psi^{am}_1$ and $\psi^{am}_2$ are approximately constants and these quantities are determined in theory \cite{TM}. In Eqs.(1-4,12-14) the system of units was used in which the reciprocal lattice constant is measured in units of $\lambda_e^{-1}$ ($\lambda_e =\hbar/mc$) and the direct lattice constant is measured in units of $\lambda_e$; this is adopted in the theory of coherent radiation and pair-production \cite{TM}. The choice of the basic vectors ${\bf{G_1}}$, ${\bf{G_2}}$, ${\bf{G_3}}$ is not unique. It is convenient to choice these vectors along axes of symmetry of the crystallographic lattice. So, for instance, let choice the vector ${\bf{G_1}}$ along the $<110>$-axis in a silicon single crystal. Then one can choice the vectors ${\bf{G_2}}$ and ${\bf{G_3}}$ along the $<001>$ and $<1 \bar{1} 0>$ axes, correspondingly. In this case the lengths of these vectors are equal to \begin{eqnarray} G_2=2\pi/a, \qquad \qquad G_3=2\sqrt{2}\pi/a, \label{15} \end{eqnarray} where the $a$ is the side of the sell. One can see from the expressions (1)-(4) that the components of the tensor $\varepsilon_{ij}$ are the functions of the two universal parameters $W_H$ and $W_V$ (if the term $\varepsilon_A$ is ignored). These parameters for silicon crystal and orientation determined by Eq.(15) are equal to \begin{eqnarray} W_H=6.183 E_\gamma \theta \sin\alpha \qquad W_V=4.372 E_\gamma \theta \cos\alpha \qquad \label{16} \end{eqnarray} where $E_\gamma$ and $\theta$ are measured in GeV and radians, correspondingly. Knowing the permittivity tensor $\varepsilon_{ij}$ one can find the refractive indices of $\gamma$-quanta \cite{MMF} \begin{equation} \tilde n^2 =(\varepsilon_{11}+\varepsilon_{22})/2 \pm \sqrt{(\varepsilon_{11}-\varepsilon_{22})^2/4 + \varepsilon_{12}\varepsilon_{21} } \, , \label{17} \end{equation} Thus two waves with different indices of refraction $\tilde n_1$ and $\tilde n_2$ can propagate in the single crystals. In general, these refractive indices are complex quantities. Besides, in general case these two waves are elliptically polarized. However in particular case when the coordinate system exists in which the tensors $\varepsilon'_{ij}$ and $\varepsilon''_{ij}$ are simultaneously diagonal (i.e., complex tensor $\varepsilon_{ij}$ is reduced to principal axes) the both waves are linearly polarized. It is obviously that the permittivity tensor is diagonal when the momentum of $\gamma$-quanta lies strictly in $({\bf{G_1}},{\bf{G_2}})$ or $({\bf{G_1}},{\bf{G_3}})$ planes (angle $\alpha =0$ or $ \pi/2$). Then the refractive indices are equal to: \begin{equation} \tilde{n_1}= \sqrt{\varepsilon_{11}} ,\,\qquad \tilde{n_2}= \sqrt{\varepsilon_{22}} \label{18} \end{equation} In this case the differences of the real and imaginary parts of refractive indices are equal to $(\alpha=0)$ : \begin{eqnarray} \mathop{\rm Re}\nolimits (\tilde{n_1}-\tilde{n_2})= {{BN\bar{\sigma}}\over {8\pi}} {\hbar \over mc} \sum_{\bf{g}} \Phi(g)\,(g_2^2-g_3^2)z_g^2F'_1(z_g) \vartheta(z_g) , \, \label{19} \\ \mathop{\rm Im}\nolimits(\tilde{n_1}-\tilde{n_2})= - {BN \bar{\sigma} \over 16} {\hbar \over mc} \sum_{\bf{g}} \Phi(g)(g_2^2 - g_3^2)F''_1(z_g) \vartheta(1-z_g) \vartheta(z_g) , \label{20} \\ z_g=1/(n_2 W_V) \label{21} \end{eqnarray} where $\vartheta$ is the Heaviside unit step function. The similar case was considered in paper \cite{C}. However, the $\gamma$-beam obtaining for experiments has some nonzero phase volume and, strictly speaking, the number of $\gamma$-quanta, which have different angles $\theta$ but fixed angle $\alpha=0$, is equal to zero. In other words, a real $\gamma$-beam have some distribution over the angle $\alpha$. Now we show in detail that this fact change noticeably the relations for calculation of the refractive indices. The components of permittivity tensor are the sum over the reciprocal lattice vectors and summation over ${\bf{g}}$ satisfies the conditions $z_g >0$ or $ 0<z_g<=1$ for the real and imaginary components, correspondingly. Let us consider the first condition (for real components). One can rewrite its in the following form: $n_2W_V +n_3W_H >0$. When $\alpha=0$ we have $W_H=0$ and get $n_2 >0$ and $n_3$ is an arbitrary integer number ($W_V \ne 0$). Now let the angle $\alpha$ is nonzero small angle. Then we get \begin{equation} n_2(G_2 \theta \cos{\alpha}) + n_3 (G_3 \theta \sin{\alpha}) >0 \label{22} \end{equation} It easy to see that set of numbers $n_2=0, \, n3= (1,2,3 ...) \mathop{\rm sign}\nolimits{\alpha}$ ($\mathop{\rm sign}\nolimits$ is the function equal to $\pm 1$ according to sign of $\alpha$) satisfies to Eq.(22). The set of obtained numbers (in case $\alpha=0$ ) is also satisfied Eq.(22). Note that is true for any small nonzero angle $\alpha$. For imaginary components the set of $n_2,\,n_3$-numbers is the same in both cases, if only the angle $\alpha$ is enough small. Now we can calculate the components of the permittivity tensor in limits $\alpha \, \rightarrow \pm \,0 $. It is clear that permittivity tensor (in pointed limit) have a diagonal form. Finally we get the following quantity of the difference of real parts of the refractive indices: \begin{equation} \mathop{\rm Re}\nolimits (\tilde{n_1}-\tilde{n_2})= {{BN\bar{\sigma}}\over {8\pi}} {\hbar \over mc} \{\sum_{\bf{g}} \Phi(g)\,(g_2^2-g_3^2)z_g^2F'_1(z_g) \vartheta(z_g) + {8 \over 15} \sum_{n_3=1}^{\infty} \Phi(n_3 G_3)(G_3n_3)^2 \} \label{23} \end{equation} Note that the left and right limits are equal in value. The difference of the imaginary parts of refractive indices is described as before by Eq(20). The add term in Eq(23) is equal to the mean-square value of the interplanar electric field (within a multiplier) \cite{MMF} . This term is independent of the $W_H$ and $W_V$ parameters. One can pointed to the similar effect in the coherent bremsstrahlung in single crystals. Let the electron beam motion in single crystal is determined by the $W_H$ and $W_V$ parameters. Then the theory predicts that the intensity of radiation of the low energy photons is small enough, when $W_H=0$ and $W_V$ is reasonably large. However the experiments show the significant exceeding of intensity of these photons relative to calculated values \cite{TM,BBC}, if the calculations is not take into account the angular divergence of the electron beam. \section{Influence of the $\gamma$-beam divergence on propagation } As it was shown in paper \cite{MMF}, in general case the $\gamma$-beam propagate in the single crystal as superposition of the two elliptically polarized waves. Birefringence is a special case of propagation of high-energy $\gamma$-quanta in single crystals, when the elliptical polarization of these waves (eigenfunctions of the problem) degenerate into linear one. The linear polarization point to the space symmetry of the problem as it was shown previously. Now we consider the important problem for the experimental observation of birefringence. We want to get the answer on the following question: Is the refractive indices and polarization states of waves, when $\gamma$-beam move near the axis of symmetry in the single crystal (in other words, when $W_H \ne 0$, but $W_H \ll W_V$), essentially changed ? With the aim of investigation of this problem we carry out calculations of the refractive indices and polarization states of waves at the small values of $W_H$. These calculations are based on papers \cite{MMF,MV1,MV2} where the general case of $\gamma$-quanta propagation in the anisotropic medium was considered. Besides, we examine only the case when the beam of $\gamma$-quanta move under a small angle $\theta$ with respect to one of the "strong" crystallographic axis (in other words when $W_H, \, W_V \, \simeq 1)$. The difference of real parts of refractive indices is more significant at these orientations in compare with the motion of beam near crystallographic planes \cite{MMF}. Figures 1 and 2 show the results of calculations of the refractive indices as functions of $W_V$ at some values of $W_H$. One can see that the variations of the refractive indices difference are insignificant as a whole when the parameter $W_H$ is within 0 - 0.01. However the peaks of curve at $W_H=0$ are spreading enough when the parameter $W_H$ rise to 0.01. The curve at $W_H=0.1$ is differ from curve at $W_H=0$ for all practically values of the $W_V$. In all calculations the Moliere form factor was employed \cite{BKF}. Figure 3 illustrates the absolute value of circular polarization, which have the normal electromagnetic waves (eigenfunctions of the problem). The circular polarization $P_c$ is small when $W_H \sim 0.01 $ and it rise to 0.5 with increasing of the parameter $W_H$ to 0.1. Nevertheless, the value of linear polarization $P_L=\sqrt{1-P_c^2}$ of the normal waves is dominant at all considered here values of $W_H$ and $W_V$. Besides, the turn of the semiaxes of polarization ellipse in plane $({\bf{G_2,G_3}})$ is take a place (see figure 4). Thus once can say that the pure birefringence in the single crystal takes a place for $\gamma$-beam with a small angle divergence (the value of this divergence one can found from relation $\delta W_H \sim 0.01$). On the other hand, the variations of polarized state of the $\gamma$-quanta propagating in single crystals at different values of $W_H$ is of more direct interest to practical goals. Figures 5,6 show the variation of the circular polarization of the 100 GeV $\gamma$-beam propagating in the silicon single crystal as a function of its thickness. The system of coordinate is chosen so that the Stokes parameter of $\gamma$-beam $\xi_3= \pm 1$ when 100 \% linear polarization lies in planes $(1 \bar{1}0)$ and $(001)$, correspondingly. We take for illustration the cases of partially polarized beam and unpolarized one at the point of entry in the single crystal. In the case of pure birefringence (see figure 5) the unpolarized $\gamma$-beam can obtain only some degree of the linear polarization on any thickness of a single crystal (i.e. $\xi_2(x) =0$). In the case when the normal electromagnetic waves is elliptically polarized the propagating unpolarized beam of $\gamma$-quanta can obtain some degree of the linear and circular polarization (see figure 6). The transformation of linear polarization to circular one ( under angle in $ \pm 45^{o}$ with respect to above-mentioned coordinate system) one can see also on these figures. The analogous curves for parameter $W_H=-0.1$ are mirror-symmetric with respect to x-coordinates. The intensity of $\gamma$-beam is decreased in $\sim \, 10^9 $ times on 100 cm of the silicon single crystal. Note that our consideration of the bierfrigence base on the theory of coherent $e^{\pm}$-pair production in single crystals \cite{TM,U}. However this theory is violated at some orientations of single crystals (in regions of so called "strong field") \cite{MP}. For silicon crystallographic planes this violation is expected at very high energy of $\gamma$-quanta $\gg 1 TeV$. \section{Conclusion} The pure birefringence of high energy $\gamma$-beam propagating near crystallographic axis (when the eigenfunctions of a problem is two linearly polarized electromagnetic wave) take a place for special (predominantly symmetric) orientations. In general case the propagating $\gamma$-beam is the superposition of two elliptically polarized waves, because of this some peculiarities in the propagation of $\gamma$-beam exist even for orientations near to the pointed symmetrical ones. For these orientations we can point on the following: \newline 1. Some noticeable degree of circular polarization of eigenfunctions exists. \newline 2. Some angle shift of the axes of the polarization ellipse takes a place also. \newline 3. The quantities of refractive indices is changed sharply for close orientations. \newline Thus it needs to take into account these effects in experimental observation of birefringence of a beam of $\gamma$-quanta with some phase volume. In addition we consider in detail the procedure of calculation of refractive indices for real $\gamma$-beams.
\section{Introduction} \label{Sec.intro} Nonclassical features of optical pulses like squeezing, sub-Poissonian statistics and entanglement have been of increasing interest for optical communication and measurement techniques \cite{DrummondPD93,SizmannA97,AndersonME97,AbramI99}. From classical optics it is well known that in fibers photons can fly in soliton-like pulses over long distances, provided that they intense enough and the fiber nonlinearity can compensate for the dispersion-assisted pulse spreading \cite{HasegawaA89,AkhmanovSA92}. Yet solitons are not rigid but very lively nonclassical species, because of the quantum noise associated with the nonlinear dynamics. Since already single-mode radiation becomes nonclassical under the influence of a Kerr nonlinearity \cite{KitagawaM86}, solitons can be expected to exhibit not only single-mode nonclassical properties but also nonclassical internal correlations. The study of quantum solitons has been also motivated by a number of possible applications \cite{DrummondPD93,SizmannA97,AbramI99}. Solitons may be used for realizing efficient and reliable sources of pulses with intensity fluctuations below the shot noise level. Nondestructive high-precision optical switching devices and logic gates may be implemented applying the concept of quantum non-demolition measurement to the collision interaction of solitons \cite{FribergSR92,SpaelterS97a,CourtyJM98}. Suppression of noise in soliton communication lines \cite{DrummondPD93,SizmannA97,AbramI99} and usage of solitons in quantum information processing \cite{AbramI99} may be other potential applications. Nonclassical properties of optical solitons have been detected in a number of experiments. In \cite{RosenbluhM91,BergmanK94} soliton squeezing is measured using homodyne detection. Photon-number squeezing of the spectrally filtered solitons is measured by direct detection in \cite{FribergSR96,SpaelterS97}. In the experiment in \cite{SchmittS98} with asymmetrical ($10/90$) fiber-loop interferometer photon-number squeezing up to $6.0\mbox{dB}$ is achieved. Recently the experimental studies have been extended to internal spectral photon-number correlations associated with narrow bandwidth soliton components \cite{SpaelterS98}. Disregarding losses, the dynamics of quantum solitons in optical fibers may be described by the operator-valued nonlinear Schr\"odinger equation, which can be solved employing the Bethe ansatz \cite{LaiY89a}. The method was used successfully to calculate field and intensity correlations in the space-time domain \cite{YaoD95a,KartnerFX96}. In practice however, the quantum nonlinear Schr\"odinger equation must be modified by supplementing it with further terms in order to include effects such as absorption, third-order dispersion, and Raman scattering. Various methods of solution have been developed and successfully applied \cite{LaiY89,SingerF92,HausHA90,LaiY90,LaiY93,% DoerrCR94,LaiY95,DrummondPD87,DrummondPD93a,CarterSJ95,% WernerMJ96,WernerMJ97,WernerMJ97a,WernerMJ97b,SchmidtE98}. Here we follow \cite{SchmidtE98} and use cumulant expansion in Gaussian approximation for studying spectral photon-number squeezing and nonclassical photon-number correlation of different spectral components of damped solitons. We consider both narrow-bandwidth spectral pulse components \cite{SpaelterS98} and pulse components with finite spectral bandwidth selected by optimally chosen square bandpass filters \cite{WernerMJ96}. The numerically obtained results illustrate the possibility to use optical solitons for generation of squeezed light and nonclassically correlated light beams also in the presence of absorption. Moreover, the results reveal that absorption can improve, under certain circumstances, the nonclassical features. The article is organized as follows. Section \ref{Sec.model} outlines the used basic concept. The relations necessary for studying nonclassical correlations are given in Sec.~\ref{Sec.CS}, and the numerical results are reported in Sec.~\ref{Sec.Results}. Finally, a summary and some concluding remarks are given in Sec.~\ref{Sec.Summary}. \section{Basic concept} \label{Sec.model} Let us first give a brief outline of the used concept for describing the propagation of a quantized optical pulse through an absorbing fiber with second order dispersion and Kerr nonlinearity (for details, see \cite{SchmidtE98} and references therein). The spatio-temporal pulse evolution in a co-moving reference frame is described in terms of slowly varying bosonic operators $\hat{a}(x,t)$ satisfying the commutation relation \begin{equation} \left[ \hat{a}(x,t),\hat{a}^{\dagger }(x^{\prime },t)\right] ={\cal A}% ^{-1}\delta (x-x^{\prime }), \label{eq.aa} \end{equation} [${\cal A}$ is the effective cross-section of the fiber]. The undamped motion is governed by the Hamiltonian \begin{equation} \hat{H}=\hbar {\cal A}\int dx\left[ \vphantom{ a^{(2)} } \textstyle\frac{1}{2}\omega ^{(2)}\left( \partial _{x}\hat{a}^{\dagger }\right) \left( \partial _{x}\hat{a}\right) +{\textstyle\frac{1}{2}}\chi \hat{a}^{\dagger }\hat{a}^{\dagger }\hat{a}\hat{a}\vphantom{ a^{(2)} } \right] , \nonumber \end{equation} where the constant ${\chi }$ is related to the third-order susceptibility $\chi^{(3)}$ as \begin{equation} {\chi }=\frac{3\chi ^{(3)}\hbar (v_{{\rm gr}} k_{{\rm c}})^{2}} {4\epsilon _{{\rm r}}^{2}\epsilon _{0}} \label{eq.chi} \end{equation} [$\epsilon _{{\rm r}}$, relative permittivity at the carrier frequency $\omega _{{\rm c}}$; $k_{{\rm c}}$, carrier wave number; $v_{{\rm gr}}$, group velocity; $\omega ^{(2)}$ $\!=$ $\!{\rm d}^2\omega (k)/{\rm d}k^2|_{k=k_{\rm c}}$]. Note that solitons can be formed either in focusing media with anomalous dispersion (\mbox{$\chi$ $\!<$ $0$}, \mbox{$\omega^{(2)}$ $\!>$ $\!0$}) or in defocusing media with normal dispersion (\mbox{$\chi$ $\!>$ $\!0$}, \mbox{$\omega^{(2)}$ $\!<$ $\!0$}). The damped motion is treated on the basis of the master equation \begin{equation} i\hbar ~\partial _{t}\hat{\rho}=[\hat{H},\hat{\rho}] +i\gamma \hat{L}\hat{\rho}, \label{eq.master} \end{equation} where \begin{eqnarray} \gamma \hat{L}\hat{\rho} &=&\gamma \hbar {\cal A}\int dx\left[ \vphantom{ a^{(2)} } N_{% {\rm th}}\left( 2\hat{a}^{\dagger }\hat{\rho}\hat{a}-\hat{\rho}\hat{a}\hat{a}% ^{\dagger }-\hat{a}\hat{a}^{\dagger }\hat{\rho}\right) \right. \nonumber\\ &&\quad \quad \quad +\left. \left( N_{{\rm th}}+1\right) \left( 2\hat{a}\hat{% \rho}\hat{a}^{\dagger }-\hat{\rho}\hat{a}^{\dagger }\hat{a}-\hat{a}^{\dagger }\hat{a}\hat{\rho}\right) \vphantom{ a^{(2)} }\right], \label{eq.L} \end{eqnarray} with $\gamma $ being the damping constant, and \begin{equation} N_{{\rm th}}=\left[ \exp \left( \frac{\omega _{{\rm c}}\hbar }{k_{{\rm B}}T}% \right) -1\right] ^{-1} \label{Nthermisch} \end{equation} ($k_{{\rm B}}$ - Boltzmann constant, $T$ - temperature). The model applies to pulses longer than $1\,$ps, otherwise the influence of additional effects such as Raman scattering and third order dispersion must be taken into account. The master equation (\ref{eq.master}) is converted, after spatial discretization, into a pseudo-Fokker-Planck equation for an $s$-parametrized multi-dimensional phase-space function, which is solved numerically using cumulant expansion in Gaussian approximation. The initial condition is realized by a multimode displaced thermal state, without internal entanglement, and it is assumed that the field expectation value corresponds to the classical fundamental soliton. In the numerical calculation, the coordinates $x$ and $t$ are scaled by the initial pulse width $x_{0}$ and the dispersion time $t_{{\rm d}}=x_{0}^{2}|\omega ^{(2)}|^{-1}$ respectively. The calculations are performed on a grid of $200$ points with discretization step $\Delta x$ $\!=$ $\!0.1\,x_0$ for an initial photon number of the pulse of $2\bar{n}$ $\!=$ $\!2|\omega ^{(2)}{\cal A}/(\chi x_{0})|$ $\!=$ $2\!\cdot\!10^{9}$ and a reservoir photon number of $N_{{\rm th}}$ $\!=$ $\!10^{-16}$. The value of $N_{{\rm th}}$ corresponds to a carrier wavelength of $\lambda _{{\rm c}}$ $\!=$ $\!2$ $\!\!\pi $ $\!\!c$ $% \!\!/\omega _{{\rm c}}$ $\!=$ $1.5\,\mu {\rm m}$ of the pulse in vacuum and a temperature of $T$ $\!=$ $\!300$K [see Eq.~(\ref{Nthermisch})]. Thus the pulse is initially prepared in a displaced thermal state that is almost a coherent state. \section{Nonclassical spectral correlations} \label{Sec.CS} In order to study spectral properties, bosonic operators in the Fourier space are introduced, \begin{equation} \label{eq.aw} \hat{a}(\omega ,t)=\frac{1}{\sqrt{2\pi }}\int_{-\infty }^{\infty }dx\ e^{i\omega x}\hat{a}(x,t). \end{equation} The operator $\hat{N}_i$ of the number of photons in a frequency interval $(\Omega_i,\Omega_i^{\prime })$, i.e., $\Omega_i$ $\!\leq$ $\!\omega$ $\!\leq$ $\!\Omega_i^{\prime}$, reads \begin{equation} \label{eq.NWW} \hat{N}_i={\cal A}\int_{\Omega_i }^{\Omega_i ^{\prime }}d\omega \ \hat{a}^{\dagger }(\omega ,t)\hat{a}(\omega ,t), \end{equation} and the photon-number variance of the beam associated with this frequency interval can be given by \begin{equation} \langle \Delta \hat{N}_{i}^{2}\rangle =\langle \hat{N}_{i}^{2}\rangle -\langle \hat{N}_{i}\rangle ^{2} =\langle :\!\Delta \hat{N}_{i}^{2}\!:\rangle +\langle \hat{N}_{i}\rangle , \label{eq.dn} \end{equation} where $:\ :$ introduces normal ordering. When the inequality \begin{equation} \langle :\!\Delta \hat{N}_{i}^{2}\!:\rangle < 0 \label{eq.dn1} \end{equation} is valid -- a condition that cannot be satisfied within the classical noise theory -- then sub-Poissonian statistics are observed, i.e., $\langle \Delta \hat{N}_{i}^{2}\rangle$ $\!<$ $\!\langle \hat{N}_{i}\rangle $. Let us now consider two nonoverlapping frequency intervals $(\Omega _{i},\Omega _{i}^{\prime })$, $i$ $\!=$ $\!1,2$. A measure of the mutual correlation of the photon-number variances in the two frequency intervals is the correlation coefficient \begin{equation} \eta_{12}=\frac{\langle \Delta\hat{N}_{1} \Delta\hat{N}_{2}\rangle } {\sqrt{\langle \Delta\hat{N}_{1}^{2}\rangle \langle \Delta\hat{N}_{2}^{2}\rangle }} \,. \label{eq.eta12} \end{equation} (note that $\langle \Delta\hat{N}_{1} \Delta\hat{N}_{2}\rangle$ $\!=$ $\!\langle :\!\Delta\hat{N}_{1} \Delta\hat{N}_{2}\!:\rangle$ for nonoverlapping frequency intervals). From the Cauchy-Schwarz inequality \begin{equation} \langle \Delta\hat{N}_{1}^{2}\rangle \langle \Delta\hat{N}_{2}^{2}\rangle - \langle \Delta\hat{N}_{1} \Delta\hat{N}_{2}\rangle^{2} \geq 0 \label{eq.CS.op} \end{equation} it follows that the absolute value of the correlation coefficient is limited to the right, \begin{equation} |\eta _{12}|\leq 1. \label{eq.eta12.1} \end{equation} Obviously, the correlation is nonclassical, if the inequality (\ref{eq.CS.op}) is violated for normally ordered quantities, i.e., \begin{equation} \langle :\!\Delta \hat{N}_{1}^{2}\!:\rangle \langle :\!\Delta \hat{N}_{2}^{2}\!:\rangle -\langle :\!\Delta \hat{N}_{1}\Delta \hat{N}_{2}\!:\rangle^{2} < 0. \label{eq.CS.dn} \end{equation} In order to give a quantitative measure of the degree of nonclassical correlation, we define the generalized correlation coefficient \begin{eqnarray} \label{eq.y12} \tilde{\eta}_{12} &=&\frac{\langle :\!\Delta \hat{N}_{1}^{2}\!:\rangle \langle :\!\Delta \hat{N}_{2}^{2}\!:\rangle -\langle :\!\Delta \hat{N}_{1}\Delta \hat{N}_{2}\!:\rangle^{2}} {\left| \langle \Delta \hat{N}_{1}^{2}\rangle \langle \Delta \hat{N}_{2}^{2}\rangle - \langle :\!\Delta \hat{N}_{1}^{2}\!:\rangle \langle :\!\Delta \hat{N}_{2}^{2}\!:\rangle \right| } \nonumber\\ &=&\frac{\eta _{11}\eta _{22}-\eta _{12}^{2}} {|1-\eta _{11}\eta _{22}|} \end{eqnarray} where \begin{equation} \eta _{ii}= \frac{\langle :\!\Delta \hat{N}_{i}^{2}\!:\rangle }{\langle \Delta \hat{N}_{i}^{2}\rangle }\leq 1. \label{eq.eta11} \end{equation} Note that $\eta _{ii}$ is negative (positive) for sub-Poissonian (super-Poissonian) statistics. Nonclassical correlation is observed if \begin{equation} \label{eq.eta111} \tilde{\eta}_{12} < 0, \end{equation} and it is easily proved that \begin{equation} \label{eq.y121} \tilde{\eta}_{12} \ge -1. \end{equation} From Eq.~(\ref{eq.y12}) it is easily seen that when \begin{equation} \langle :\!\Delta \hat{N}_{1}^{2}\!:\rangle \langle :\!\Delta \hat{N}_{2}^{2}\!:\rangle <0, \label{eq.y12.1} \end{equation} then the correlation of the photon-number variances is nonclassical. Note that for Gaussian quantum states the criterion (\ref{eq.CS.dn}) is closely related to that used in \cite{SpaelterS98}. Similarly, the mutual correlation of the numbers of photons in the two frequency intervals can be considered, \begin{equation} \tau_{12}=\frac{\langle \hat{N}_{1} \hat{N}_{2}\rangle } {\sqrt{\langle \hat{N}_{1}^{2}\rangle \langle \hat{N}_{2}^{2}\rangle }}\, \label{eq.eta121} \end{equation} ($|\tau_{12}|$ $\!\leq 1$). Nonclassical correlation is realized if \begin{equation} \langle :\!\hat{N}_{1}^{2}\!:\rangle \langle :\!\hat{N}_{2}^{2}\!: \rangle - \langle :\!\hat{N}_{1} \hat{N}_{2}\!: \rangle ^{2} < 0 , \label{eq.CS.op1} \end{equation} or equivalently \begin{equation} \label{eq.CS.op2} \tilde{\tau}_{12} < 0 , \end{equation} with \begin{equation} \tilde{\tau}_{12} =\frac{\langle :\!\hat{N}_{1}^{2}\!:\rangle \langle :\!\hat{N}_{2}^{2}:\rangle -\langle :\!\hat{N}_{1}\hat{N}_{2}\!:\rangle ^{2}} {\langle \hat{N}_{1}^{2}\rangle \langle \hat{N}_{2}^{2}\rangle -\langle :\!\hat{N}_{1}^{2}\!:\rangle \langle :\!\hat{N}_{2}^{2}\!:\rangle } \label{eq.Y12} \end{equation} ($\tilde{\tau}_{12}$ $\!\ge$ $\!-1$). Note that for the photon numbers an inequality analogous to (\ref{eq.y12.1}) is not valid, since the average of the normally ordered square of a photon number cannot be negative. The spectral photon-number statistics can be determined using a setup shown in Fig.~\ref{fig.setup}. The scheme in Fig.~\ref{fig.setup}$(a)$ is used in \cite{FribergSR96} (and with slightly modified detection in \cite{SpaelterS97}) for measuring spectral photon-number squeezing. The scheme can also be used to reconstruct correlations between different spectral components from the measured data \cite{SpaelterS98}. However with regard to direct correlation measurement, a scheme of the type shown in Fig.~\ref{fig.setup}$(b)$ may be better suited for that purpose. Omitting the detectors, the schemes can also be used for selecting from the original beam partial beams that are highly nonclassical. \section{Results} \label{Sec.Results} \subsection{Narrow bandwidth intervals} \label{narrow} In Fig.~\ref{fig.nn0t} the temporal evolution of the coefficient $\eta _{11}$, Eq.(\ref{eq.eta11}), is plotted for a narrow bandwidth mid-interval $(\Omega _{1},\Omega _{1}^{\prime })$ $\!=$ $(-\Delta \omega,\Delta \omega)$, $\Delta\omega/\omega_0$ $\!\ll$ $\!1$, and various values of the damping parameter $\gamma$ (here and in the following frequencies are scaled by $\omega_0$ $\!=$ $x_0^{-1}$ \cite{FreqNormalization}). The strongest sub-Poissonian effect is observed for $\gamma t_{{\rm d}}$ $\!=$ $\!0.03$ at $t$ $\!\sim$ $\!10t_{{\rm d}}$. Surprisingly, the sub-Poissonian statistics of the narrow bandwidth mid-component can be stronger for a damped soliton than for the undamped one \cite{SchmidtE98}. The temporal evolution of the coefficient $\eta _{11}$ as a function of the frequency $\omega$ [for a narrow bandwidth interval $(\Omega _{1},\Omega _{1}^{\prime })$ $\!=$ \mbox{$\!(\omega$ $\!-$ $\!\Delta \omega,$ $\!\omega$ $\!+$ $\!\Delta\omega$)}] is plotted in Fig.~\ref{fig.etaxt} for $(a)$ $\gamma$ $\!=$ $\!0$ and $(b)$ $\gamma t_{{\rm d}}$ $\!=$ $\!0.03$. From the figure it is seen that the interference pattern which is typically obtained in the limit of vanishing absorption [Fig.~\ref{fig.etaxt}$(a)$] is not observed for an absorbing fiber [Fig.~\ref{fig.etaxt}$(b)$]. Note that in the limit of vanishing absorption the results obtained here are in good agreement with those obtained by stochastic simulations within the frame of the positive $P$ representation \cite{WernerMJ97b}. Disregarding absorption and solving the cumulant evolution equations given in \cite{SchmidtE98} in a linearization approximation, the solution can be given by eigenfunction expansion, the expansion coefficients being determined by the chosen initial condition. The result of superposition then corresponds to an interference pattern like that in Fig.~\ref{fig.etaxt}$(a)$. Obviously, the (phase-sensitive) terms that are superimposed do not respond uniformly to absorption, so that the internal coherences responsible for the interference pattern can be destroyed at least in part. In particular, near the center of the spectrum \mbox{($\omega$ $\!\to$ $\!0$)} the super-Poissonian peaks are fully suppressed and in place of them sub-Poissonian statistics is observed. Accordingly, the super-Poissonian side-band formation that appears with increasing frequency is more uniform for an absorbing fiber than for a nonabsorbing one. To get an insight into the correlation between spectral components at different frequencies $\omega_1$ and $\omega_2$ [for narrow bandwidth intervals $(\Omega _{i},\Omega _{i}^{\prime })$ $\!=$ \mbox{$\!(\omega_i$ $\!-$ $\!\Delta \omega,$ $\!\omega_i$ $\!+$ $\!\Delta\omega$)}, $i$ $\!=1,2$], we have plotted the correlation coefficients $\eta _{12}$ [Eq.~(\ref{eq.eta12}), Fig.~\ref{fig.etaww}], $\tilde\eta_{12}$ [Eq.~(\ref{eq.y12}), Fig.~\ref{fig.cs02dnww}], and $\tilde\tau_{12}$ [Eq.~(\ref{eq.Y12}), Fig.~\ref{fig.cs02ww}] as functions of $\omega_1$ and $\omega_2$ for \mbox{$\gamma$ $\!=$ $\!0$} and \mbox{$\gamma t_{\rm d}$ $\!=$ $\!0.03$}, restricting our attention to the two characteristic propagation times \mbox{$t$ $\!=$ $\!5\,t_{\rm d}$} and \mbox{$t$ $\!=$ $\!10\,t_{\rm d}$} (cf. Fig.~\ref{fig.etaxt}). Figures \ref{fig.etaww}$(a)$ and \ref{fig.etaww}$(b)$ reveal that the super-Poissonian side-band components observed at the propagation time \mbox{$t$ $\!=$ $\!5\, t_{\rm d}$} (Fig.~\ref{fig.etaxt}) are strongly positively correlated with respect to the photon-number variance, whereas relatively strong negative correlation between the super-Poissonian side-band components and the sub-Poissonian mid-components is observed -- effects which can be found for both a nonabsorbing fiber, Fig.~\ref{fig.etaww}$(a)$, and an absorbing fiber, Fig.~\ref{fig.etaww}$(b)$. The notably reduced correlation observed at the propagation time \mbox{$t$ $\!=$ $\!10\, t_{\rm d}$} for a nonabsorbing fiber, Fig.~\ref {fig.etaww}$(c)$, may be viewed as an interference effect in a similar sense as mentioned above. From that it can be understood that absorption can enhance correlation -- a surprising effect which can be seen from Fig.~\ref {fig.etaww}$(d)$. From Fig.~\ref{fig.cs02dnww} it is seen that nonclassical correlation of the photon-number variance appears between sideband super-Poissonian components and -- in agreement with the condition (\ref{eq.y12.1}) -- between the sub-Poissonian mid-component and the super-Poissonian sideband components. The relatively strong nonclassical correlation \mbox{$\tilde\eta_{12}$ $\!\sim$ $\!-7.3\!\cdot\!10^{-3}$} observed for a nonabsorbing fiber at the propagation time $t$ $\!=$ $\!5\,t_{{\rm d}}$ [Fig.~\ref{fig.cs02dnww}$(a)$] reduces to \mbox{$\tilde\eta_{12}$ $\!\sim$ $\!-8.6\!\cdot\!10^{-4}$} at $t$ $\!=$ $\!10\,t_{{\rm d}}$ [Fig.~\ref{fig.cs02dnww}$(c)$]. The behavior of $\tilde\eta_{12}$ for an absorbing fiber may be quite different, as is seen from comparison of Figs.~\ref{fig.cs02dnww}$(a)$ and \ref{fig.cs02dnww}$(c)$ with Figs.~\ref{fig.cs02dnww}$(b)$ and \ref{fig.cs02dnww}$(d)$, respectively. Whereas at $t$ $\!=$ $\!5\,t_{{\rm d}}$ the effect of nonclassical correlation is weaker, \mbox{$\tilde\eta_{12}$ $\!\sim$ $\! -3\!\cdot\!10^{-3}$} [Fig.~\ref{fig.cs02dnww}$(b)$], a stronger effect is observed at $t$ $\!=$ $\!10\,t_{{\rm d}}$, \mbox{$\tilde\eta_{12}$ $\!\sim$ $\! -4.8\!\cdot\!10^{-3}$} [Fig.~\ref{fig.cs02dnww}$(d)$]. The latter reflects the enhanced correlation between the mid-component and the sideband components as shown in in Fig.~\ref{fig.etaww}$(d)$. We are again left with the surprising fact that absorption can enhance nonclassical behavior. Figure \ref{fig.cs02ww} shows that the photon-number correlation can be quite different from the correlation of the photon-number variance. Contrary to the photon-number variance, there is no nonclassical correlation of the photon numbers of the sub-Poissonian mid-component and the super-Poissonian sideband components, since for the photon number an inequality of the type (\ref{eq.y12.1}) cannot be valid. Nevertheless, the minimum value \mbox{$\tilde\tau_{12}$ $\!\sim$ $\! -0.02$} obtained for a nonabsorbing fiber at the propagation time \mbox{$t$ $\!=$ $\!5\,t_{{\rm d}}$} [Fig.~\ref{fig.cs02ww}$(a)$] is smaller than the minimum value \mbox{$\tilde\eta_{12}$ $\!\sim $ $\!-7.3\!\cdot\!10^{-3}$} [Fig.~\ref{fig.cs02dnww}$(a)$], and hence a stronger nonclassical correlation is observed for the photon number than the photon-number variance. Moreover, the effect of nonclassical correlation observed at the propagation time \mbox{$t$ $\!=$ $\!10\,t_{{\rm d}}$} is weaker than that at \mbox{$t$ $\!=$ $\!5\,t_{{\rm d}}$} for both a nonabsorbing fiber [Fig.~\ref{fig.cs02ww}$(c)$] and an absorbing one [Fig.~\ref{fig.cs02ww}$(d)$] [contrary to the correlation of the photon-number variance, Fig.~\ref{fig.cs02dnww}$(b)$ and $(d)$]. \subsection{Finite bandwidth intervals} \label{finite} Let us return to the coefficient $\eta _{11}$. Allowing for finite frequency windows, we consider a frequency interval $(\Omega _{1},\Omega _{1}^{\prime })$ $\!=$ $\!(-\Omega,\Omega)$ [for the scheme, see Fig.~\ref{fig.setup}$(a)$] and optimize $\Omega$ such that the sub-Poissonian components dominates the signal and $\eta _{11}$ becomes minimal for all times (Fig.~\ref{fig.eta11x3}). From Fig.~\ref{fig.eta11x3}$(d)$ it is seen that for a nonabsorbing fiber the absolute minimum of $\eta _{11}$ that is attainable in this way is $\eta _{11}$ $\!\sim$ $\!-3.5$ ($t$ $\!\sim$ $\! 4.5\,t_{{\rm d}}$), which corresponds to a Fano factor of \mbox{$F_1$ $\!=$ $\!\langle\Delta\hat{N}_1^2\rangle/ \langle\hat{N}_1\rangle$} $\!=$ $\!(1$ $\!-$ $\!\eta_{11})^{-1}$ $\!\sim$ $\!0.23$ or $\sim 6.4\,\mbox{dB}$ squeezing (cf. \cite{WernerMJ96,MecozziA97}). In that case $\sim$ $\!80\%$ of the photons of the initial, full pulse contribute to the filtered pulse [Fig.~\ref{fig.eta11x3}$(c)$]. It is further seen that the oscillating behavior of $\eta _{11}$ as a function of the propagation time is damped out owing to absorption such that for certain propagation times stronger sub-Poissonian statistics can be observed for an absorbing fiber than a nonabsorbing one. This effect is of course a consequence of the behavior of the narrow bandwidth components as addressed in Sec.~\ref{narrow}. The observed discrepancy between $\Omega$ [Fig.~\ref{fig.eta11x3}$(a)$] and $\Omega _{0}$ [Fig.~\ref{fig.eta11x3}$(b)$, for the definition of $\Omega _{0}$, see Fig.~\ref{fig.etaxt}], e.g., for $\gamma t_{\rm d}$ $\!=$ $\!0\ldots 0.005$ at $t$ $\!\gtrsim$ $\!9\,t_{\rm d}$, obviously results from the correlation between different frequency components (cf. \cite{FribergSR96,WernerMJ96}). To obtain strong nonclassical correlation between different beams, we first consider two (with respect to the center of the spectrum) symmetric frequency windows $(\Omega_{1},\Omega _{1}^{\prime })$ and $(\Omega_{2},\Omega'_{2})$ $\!=$ $\!(-\Omega'_{1},-\Omega _{1})$ [for the scheme, see Fig.~\ref{fig.setup}$(b)$] and optimize $\Omega_{1}$ and $\Omega_{1}'$ such that $\tilde\eta_{12}$ (Fig.~\ref{fig.y12nw}) and $\tilde\tau_{12}$ (Fig.~\ref{fig.yy12nw}) become minimal. From inspection of Fig.~\ref{fig.y12nw}$(d)$ the absolute minimum of $\tilde\eta_{12}$ is realized for a nonabsorbing fiber ($\tilde\eta_{12}$ $\!\sim$ $\! -0.48$ at $t$ $\!\sim$ $\! 4.5\,t_{{\rm d}}$). The effect of absorption is again seen to damp out the oscillations of $\tilde\eta_{12}$ (as a function of the propagation time) such that at certain propagation times [$\gamma t_{\rm d}$ $\!=$ $\!0.01\ldots 0.02$, $t/t_{\rm d}$ $\!=$ $\!6\ldots 14$ in Fig.~\ref{fig.y12nw}$(d)$] stronger nonclassical correlation can be achieved for an absorbing fiber than a nonabsorbing one. On the contrary, Fig.~\ref{fig.yy12nw}$(d)$ reveals that dissipation always reduces the nonclassical photon-number correlation, i.e., for any propagation time the lower bound of the coefficient $\tilde\tau_{12}$ is realized in the limit of vanishing absorption. Note that the strongest nonclassical photon-number correlation ($\tilde\tau_{12}$ $\!=$ $\!-0.32$ at $t$ $\!\sim$ $\!3.5\,t_{{\rm d}}$) is weaker than the strongest nonclassical correlation of the photon-number variance. For weak absorption ($\gamma t_{\rm d}$ $\!=$ $\!0\ldots 0.005$) from Figs.~\ref{fig.y12nw}$(a)$ and $(b)$ a (quasi-)periodic change of the frequency interval $(\Omega_{1}, \Omega _{1}^{\prime })$ is seen, in agreement with the results in Sec.~\ref{narrow}. At the early stage of propagation the frequency interval is essentially determined by the nonclassically correlated super-Poissonian sideband components. Later [at $t$ $\!\sim$ $\!3 \,t_{{\rm d}}$ in Figs.~\ref{fig.y12nw}$(a)$ and $(b)$] it shifts towards the mid-frequency, which indicates the increasing weight of the sub-Poissonian mid-components. The shift back to the sideband components obviously results from the \mbox{(quasi-)}periodic formation of super-Poissonian (and sub-Poissonian) components in the center of the spectrum [cf. Fig.~\ref{fig.etaxt}$(a)$]. The fraction of the number of photons in each beam relative to the number of photons in the initial, full beam changes from less then $5\%$ for sideband frequency windows to about of $40\%$ for near mid-frequency windows [Fig.~\ref{fig.y12nw}$(c)$]. With increasing absorption only the first shift of the frequency window from the sideband to the central part is observed. In contrast to $\tilde\eta_{12}$, minimization of $\tilde\tau_{12}$ always requires sideband frequency windows [Figs.~\ref{fig.yy12nw}$(a)$ and $(b)$], where each beam contains less than $10\%$ of the initial number of pulse photons [Fig.~\ref{fig.yy12nw}$(c)$]. Whereas symmetric windows are expected to be best suited to realization of minimal $\tilde\tau_{12}$, from Fig.~\ref{fig.cs02dnww} it is suggested that asymmetric windows may be more suited to realization of minimal $\tilde\eta_{12}$. This is fully confirmed by the calculation. The symmetric windows shown in Fig.~\protect\ref{fig.yy12nw} indeed yield the smallest value of $\tilde\tau_{12}$. A comparison of Fig.~\ref{fig.y12nw}$(d)$ with Fig.~\ref{fig.et12} reveals that $\tilde\eta_{12}$ can be reduced if asymmetric windows are used. In particular, it is seen that the absolute minimum of $\tilde\eta_{12}$ observed for a nonabsorbing fiber can be reduced to $\tilde\eta_{12}$ $\!\sim$ $\! -0.71$ ($t$ $\!\sim$ $\!5.5t_{{\rm d}}$). A more detailed analysis is given in Fig.~\ref{fig.w4_15x3}. It is seen that only at a early stage of propagation symmetric and asymmetric frequency windows yield equal values of $\tilde\eta_{12}$. Obviously, asymmetric frequency windows take better account of the nonclassical correlation between sub-Poissonian mid-components and super-Poissonian sideband components in order to reduce $\tilde\eta_{12}$. The price to be paid are the unequal photon numbers of the two beams, since the number of photons in the spectral interval around the center is substantially larger than that in the sideband interval, as is seen from the figure. Hence for generation of (with respect to the photon-number variance) nonclassically correlated beams, the frequency windows should be chosen such that an optimum between nonclassical correlation and beam intensities is observed. \section{Summary and concluding remarks} \label{Sec.Summary} We have studied the internal quantum statistics of fundamental solitons in absorbing Kerr media, applying multimode cumulant-expansion techniques and solving the resulting evolution equations numerically in Gaussian approximation. In particular, we have calculated the temporal evolution of the photon-number variance, its correlation, and the photon-number correlation for various frequency windows. The formation of super-Poissonian sideband components with nonclassically correlated photon numbers and nonclassically correlated photon-number variances may be regarded as being a typical signature of the quantum nature of a soliton pulse. Since a soliton pulse is a highly involved multimode field, interference effects can essentially determine its properties. It is worth noting that absorption can damp out interferences that are destructive with respect to nonclassical features, such as squeezing and the nonclassical correlations mentioned, so that absorption surprisingly improves, under certain conditions, these nonclassical effects. The calculations show that for a nonabsorbing fiber propagation distances of $3.5\ldots 5.5$ dispersion lengths are best suited for detecting the nonclassical features. Destructive interferences are observed at distances of $8\dots 12$ dispersion lengths. At these distances the best values for photon-number squeezing and nonclassical correlation of the photon-number variance are achieved for an absorbing fiber ($\gamma t_{{\rm d}}$ $\!\sim$ $\!0.02$). Using appropriately chosen spectral windows, a soliton pulse can serve as a source of nonclassically correlated light beams. We have considered both symmetric and asymmetric windows and optimized them such that the filtered beams are maximally nonclassically correlated with regard to the photon number and the photon-number variance. It should be pointed out that in practice a number of effects such as Raman scattering and third-order dispersion should be included in a refined theoretical model. Finally, inclusion in the calculation of non-Gaussian effects for weak absorption has been a challenge. \acknowledgments This work was supported by the Deutsche Forschungsgemeinschaft. We are grateful to F.~K\"onig, G.~Leuchs, and A.~Sizmann for valuable discussions. \widetext \begin{appendix} \narrowtext \widetext \end{appendix} \vspace*{2cm} \narrowtext \vspace*{-2.5cm}
\section{Introduction} It is widely believed that the rapid variability frequently observed in blazars (for reviews see Ulrich, et al. 1997; Wagner, 1997) is linked to dissipation in relativistic jets. The characteristically short variability time scales - down to a few minutes at optical and x-ray wavelengths (e.g., Wagner 1997), and several hours at gamma-ray energies (Mattox, J., et al. 1997; Aharonian et al. 1998) - indicate that the source is compact and strongly beamed, supporting this view. Despite the apparent complexity of the temporal behavior exhibited by blazars, ongoing observational efforts have revealed some trends. Firstly, the high-and low-energy emission appears to be well correlated. Correlations between optical/UV and gamma-ray emission have been observed in many blazars (Wagner, 1997, and references therein), and between the X-ray and TeV emission in the TeV BL Lac objects (Macomb et al. 1995; Buckley et al. 1996; Aharonian et al. 1998). There is also evidence for a correlated activity at radio and gamma-ray bands in some cases (e.g., 3C279; Wehrle et al. 1998). Secondly, the variation of the synchrotron emission below the spectral turnover has the tendency to be slower and later at longer wavelengths (with the exception of radio IDV), suggesting that the low frequency (radio-to-IR) emission is self-absorbed at the early stages of the outburst. In spite of these trends, it seems that blazars can exhibit a variety of temporal behaviors, and that even in individual sources the variability pattern can change, as exemplified by observations of the BL Lac object PKS 2155-304 that revealed different variability patterns during the 1991 (e.g., Brinkmann et al. 1994; Edelson et al. 1995) and 1994 (Urry et al. 1997) campaigns. The change in temporal behavior can be ascribed to the different physical conditions that exist in the source during each episode, and it is, therefore, desirable to examine the dependence of the temporal characteristics of the emission on the physical parameters of the source. In this paper we study the time dependent synchrotron emission from relativistic jets in the framework of the radiative front model developed previously to study gamma-ray flares (Levinson, 1998; hereafter Paper I). The model is extended to encompass synchrotron emission and employed to calculate the temporal evolution of the synchrotron and ERC emission radiated by the front, with an attempt to elucidate the dependence of the variability pattern on the model parameters. As shown below, the model can successfully account for some of the trends mentioned above, and predicts some additional features. In \S 2 we give an outline of the model and the necessary extensions. In \S 3 we derive the equation governing the evolution of the synchrotron flux in the front, and explain how it is incorporated into the numerical model. We then go on to describe the numerical results. The observational consequences are discussed in \S 4. \section{Description of the model} The model presented in Paper I assumes that the variable emission seen in blazars originates inside dissipative fronts that are produced by overtaking collisions of highly magnetized, relativistic outflows, and consists of a pair of shocks and a contact discontinuity (Romanova \& Lovelace, 1997; Levinson \& Van Putten, 1997). The front propagates with a velocity intermediate between that of the slow and fast fluid slabs, dissipates a fraction of the outflow energy, and cools radiatively and adiabatically. A fraction of the dissipation power is taped for the acceleration of electrons to high energies, and the rest for the heating of the bulk plasma in the front. The acceleration time of non-thermal electrons has been assumed to be short compared with the cooling and the corresponding light crossing times. Further, the fluid parameters and the fraction of dissipation power that is taped for electron acceleration are assumed to have no explicit time dependence (but do depend on time implicitly through the dynamics of the front). As a consequence, the characteristics of the transient emission produced by this model reflect essentially the dynamics of the system, as well as the intensity of external radiation, rather than explicit variations of the outflow parameters and/or electron acceleration rate, as, e.g., in the one-zone, homogeneous model by Mastichiadis \& Kirk (1997). In Paper I we focused on the production of gamma-ray flares in the regime where ERC emission (e.g., Dermer \& Schlickheiser 1993; Blandford \& Levinson 1995) dominates over SSC emission (Bloom \& Marscher 1993). The dynamics of the system and the time evolution of the flux radiated by the front have been calculated in a self-consistent manner, by numerically solving the MHD equations governing the front structure coupled to the kinetic equations describing the angle averaged pair and gamma-ray distribution functions. The solutions depend on four key parameters: i) the dissipation rate of magnetic energy, ii) the maximum injection energy of electrons, denoted by $E_{emax}$, which was taken to be fixed in the observer frame for simplicity, iii) the fraction of dissipation energy that is injected in the form of non-thermal electrons, and iv) the ratio of the thickness of ejected fluid slab and the gradient length scale of background radiation intensity. The latter parameter determines essentially whether the shape and timescale of the flare are related to the radial variation of ambient radiation intensity, in which case the light curve is asymmetric with a rapid rise and a longer decay, or to the shock travel time across the fast fluid slab, in which case the decay time is of order the cooling time and is typically much shorter than in the case of infinite length outbursts. As shown in Paper I, the temporal evolution of the gamma-ray flux is insensitive to the injected electron spectrum provided that electron acceleration is efficient, and that $E_{emax}$ exceeds the corresponding gamma-spheric energy at any given radius along the course of the front. The evolution of the synchrotron spectrum, on the other hand, is expected to depend on variations of $E_{emax}$, particularly if the front is optically thin at the corresponding frequencies (cf. Levinson 1996). We shall, therefore, relax the assumption that $E_{emax}$ is fixed. In general, the evolution of the injected electron distribution is governed by the acceleration process, and a self-consistent treatment of particle acceleration is required in order to follow the development of $E_{emax}$. Such a treatment is beyond the scope of our paper. Here we settle for a simple prescription where the injected spectrum is taken to be a power law with an exponential cutoff above $E_{emax}(t)$, where \begin{equation} E_{emax}(t)=E_o (r/r_o)^{b}, \label{eq:Emx} \end{equation} with $r(t)=r_o+\beta_c ct$. One important feature of the model is a positive radiative feedback that affects the emitted spectrum considerably. The point is that as a result of radiative losses the front decelerates and its expansion rate decreases until the peak of the radiated power is reached. The deceleration of the front leads, in turn, to enhanced dissipation rate of the bulk flow energy, since the relative velocity between the fast fluid and the reverse shock increases. After peak emission is reached, the front begins to reaccelerate, ultimately reaching its initial speed and structure. The consequences of this radiative feedback for the variability have been discussed in detail in Paper I. Another consequence is that the variability should depend strongly on the orientation of the source, owing to the change in beaming factor during the deceleration phase. The analysis presented in Paper I cannot account for such orientation effects, as it only treats the angle averaged flux. The model does describe rather well, however, the variability that would be seen by an observer at sufficiently small viewing angle. A study of orientation effects will be presented elsewhere (Eldar \& Levinson, in preparation). \section{Synchrotron flares} In this section we extend the analysis to incorporate synchrotron emission, and apply the results to study the relation between the low-and high-energy emission in ERC dominated blazars. In order to do so, we augment the set of equations introduced in Paper I by an equation governing the time change of the synchrotron intensity (see eq. [\ref{eq:transfer}] below). We also include additional loss term for the electrons accounting for synchrotron cooling. The equations are then integrated in the injection frame (the rest frame of the boundary from which the fluid is expelled), as described in Paper I. Now, the synchrotron opacity depends on the electron density and magnetic field inside the front and, consequently, on the transverse expansion of the front. Let $c_{s\perp}$ denotes the expansion speed in the transverse direction, $d$ the cross-sectional radius, and $A=\pi d^2$ the cross-sectional area. The expansion speed $c_{s\perp}$ may depend on external pressure and magnetic fields, or the density of ambient gas if inertial effects are important, and should not be the same as the sound speed in the radial direction in general. We suppose that initially $d_o= \psi r_o$, where $\psi$ is the jet opening angle, and $r_o$ is the radius of formation of the front (Paper I). Then $d=d_o+c_{s\perp}t= \psi[r_o+(c_{s\perp}/\psi)t]$. For illustration we take $ (c_{s\perp}/\psi)$ to be equal to the front velocity, $c\beta_c$. We then obtain $A(t)\propto r^2(t)$, where $r(t)=(r_o+c\beta_c t)$. Let $N_e$ be the total number of electrons in the front. The corresponding number density is then given by $n_e=N_e/(\Delta X A)$, where $\Delta X$ is the axial length of the front. The magnetic field is assumed to decline like \begin{equation} B=B_o [r(t)/r_o]^{-p}. \label{eq:B} \end{equation} \subsection{Transfer equation} We define $I_{\nu}(t,\mu)$ to be the unpolarized synchrotron intensity inside the front, as measured in the injection frame. The equation governing the time evolution of $I_{\nu}$ can be derived in the manner described in Paper I. One then obtains, \begin{equation} \frac{\partial}{\partial x^0}I_{\nu}=j-[(\mu-\beta_{s-})/ (\Delta X)+\kappa] I_{\nu}, \label{eq:transfer} \end{equation} where $j(\nu,\mu,t)$ and $\kappa(\nu,\mu,t)$ are the emission and absorption coefficients, $x^0=ct$, $\beta_{s-}$ is the velocity of the reverse shock, $\cos^{-1}\mu$ is the angle between the direction of a photon and the front velocity, and $\Delta X$ is the axial length of the front. The term proportional to $(\mu-\beta_{s-})$ accounts for the change in the intensity due to the combined effects of front expansion and escape of synchrotron photons from the front (see Paper I for details). For convenience, we compute the emission and absorption coefficients in the rest frame of the front (quantities measured in the front frame are henceforth denoted by prime), and then transform them to the injection frame. The coefficients in the two frames are related through (Rybicki \& Lightman 1979), \begin{eqnarray} j(\nu,\mu)=\left(\frac{\nu}{\nu'}\right)^2j'(\nu',\mu'), \\ \kappa(\nu,\mu)=\frac{\nu'}{\nu}\kappa'(\nu',\mu'), \end{eqnarray} with $\nu={\cal D}\nu'$, ${\cal D}=[\Gamma_c(1-\beta_c\mu)]^{-1}$ being the Doppler factor, where $\Gamma_c$ is the bulk Lorentz factor of the front. The emission and absorption coefficients are given in the front frame by, \begin{equation} j'=\frac{\sqrt{3}e^3}{4\pi mc^2}B(t)\sin\phi'\int{n'_e({\cal E}',t) F(x') d{\cal E}'}, \end{equation} \begin{equation} \kappa'=\frac{\sqrt{3}e^3}{8\pi m \nu'^2}B(t)\sin\phi'\int{{\cal E}'^2 \frac{d}{d{\cal E}'}\left[\frac{n'_e}{{\cal E}'^2}\right] F(x') d{\cal E}'}. \label{eq:kp} \end{equation} Here ${\cal E}'$ is the electron energy with respect to the front frame, $n'_e$ is the corresponding electron number density per unit energy, $\phi'$ is the angle between the line of sight and the magnetic field in the front frame, $x'=\nu'/\nu'_c$; $\nu'_c=(3e/4\pi mc)B(t)\sin\phi' ({\cal E}'/mc^2)^2$, and $F(x)=x\int_{x}^{\infty}K_{5/3}(\chi)d\chi$. We emphasize that $B(t)$, $\Gamma_c(t)$, and $\beta_{s-}(t)$ are not given explicitly but rather calculated self-consistently by integrating the front equations. Now, as mentioned above, the integration of the electron kinetic equation is carried out in the injection frame. This yields $n_e( t,{\cal E})$. Thus, in order to obtain $n'_e(t,{\cal E}')$ we need to transform $n_e$ into the front frame. The appropriate transformation is derived in the Appendix under the assumption that the comoving electron distribution is isotropic. As a check on the above equation, consider the case of a radiating blob, for which $\beta_{s+}=\beta_{s-}=\beta_{c}$. It can be readily shown that the steady-state solution of eq. (\ref{eq:transfer}) reduces, in the optically thin limit, to $I=j\Delta X/(\mu-\beta_c).$ On substituting $\Delta X'=\Gamma_c\Delta X$ and $\mu'=(\mu-\beta_c)/ (1-\beta_c\mu)$ in the latter expression, we recover the well known result: $I={\cal D}^3(j'\Delta X'/\mu') ={\cal D}^3 I'$. Note that $\Delta X'$ and $\mu'$ are respectively the blob thickness and the cosine of the angle between the blob velocity and the direction of a photon, as measured in the blob frame. We shall proceed by assuming that the magnetic field in the front is tangled and has no net direction. The emission and absorption coefficients can then be averaged over $\phi'$. Since we do not consider polarization and orientation effects, we anticipate the results not to depend strongly on this assumption. We further average eq. (\ref{eq:transfer}) over the viewing angle $\mu$. \subsection{Cooling rates} The synchrotron cooling time can be expressed in terms of the front parameters as \begin{equation} \frac{ct_{syn}}{r_o}=0.17\left(\frac{L_{j46}}{r_{o16}}\right)^{-1} \Gamma_{A-}^2(n_f/n_{-})U^2_{Af}({\cal E}/mc^2)^{-1}. \end{equation} Here $n_f$, $n_{-}$ are the electron densities inside the front and in the fluid exterior to the front, respectively, $U_{Af}$ is the Alfv\'en 4-velocity with respect to front frame, $L_{j46}$ is the power carried by the fast outflow in units of $10^{46}$ erg s$^{-1}$, $\Gamma_{A-}$ is the Lorentz factor associated with the Alfv\'en 4-velocity of the exterior fluid, and $r_{o16}$ is the radius of shock formation $r_o$ in units of $10^{16}$ cm. For magnetized jets $\Gamma_{A-}>>1$. The product ($n_f/n_{-})U_{Af}$ changes with time due to the radiative feedback (Paper I), and is typically of order 30 or so in the presence of rapid magnetic field dissipation. Thus, the synchrotron cooling time at most electron energies is much shorter than the dynamical time $r_o/c$. The ratio of synchrotron and Compton cooling rates equals the ratio of comoving energy densities of magnetic field and external radiation field, and is given by \begin{equation} \frac{t_{IC}}{t_{syn}}\simeq \frac{20}{\Gamma_{A-}^2\Gamma_c^2} \frac{L_{j46}}{(\epsilon L_{s})_{45}}(n_f/n_{-})U^2_{Af} \label{eq:cool} \end{equation} with $(\epsilon L_{s})_{45}$ being the fraction of soft-photon luminosity that is scattered across the jet in units of $10^{45}$ erg s$^{-1}$. This ratio is independent of radius when $p=1$ in eq. (\ref{eq:B}) (provided the soft photon intensity declines as $r^{_2}$). Nevertheless, enhanced compression during the radiative phase renders it time dependent, as shown in fig. 1. For the parameters adopted in fig. 1 electron cooling is dominated by ERC emission as long as $L_j<10(\epsilon L_{s})$. \subsection{Numerical results} The numerical model presented in Paper I has been extended to include equations (\ref{eq:transfer})-(\ref{eq:kp}). The entire set of equations has then been integrated, starting from $r_o$ where $I_{\nu}=0$, and where the front structure is taken to be that of an adiabatic front (see Paper I for a detailed discussion concerning the initial conditions). In the following examples the Lorentz factors of the fluids ahead and behind the front and the rest frame Alfven 4-velocity have been chosen to be respectively 5, 20, and 10, as in Paper I. A rapid magnetic field dissipation with the same decay constant as in Paper I ($\alpha_b=0.5$), and a background radiation field with the same intensity have been invoked. We examine first the dependence of the variability pattern on the thickness of expelled fluid slab, $d$, in the case in which the maximum injection energy, $E_{emax}$, is time independent. As shown below, both the shape of the light curve and the spectral evolution depend on $d$. Quite generally we find that the ejection of slabs of thickness $d>>r_o$ leads to the production of asymmetric flares with a fast, exponential rise and a much slower, power law decline, as shown in the {\it bottom right} panel of fig. 2, where a sample light curves of the optically thin emission, computed for different values of $d/r_o$ is exhibited . The decay time and profile in this case are determined predominantly by the decline of the density and magnetic field inside the front due to its transverse expansion. When the expelled slab is sufficiently thin, specifically, when $d/r_o$ is smaller than approximately $(1-\beta_c)$, $\beta_c$ being the average front velocity, shock crossing of the fluid slab followed by a large drop of the energy dissipation and consequent particle acceleration rates occurs on a timescale much shorter than the time change of the front parameters due to expansion. As a result, the radiated flux decay on the cooling time scale (provided it is short enough), leading to a roughly symmetric light curve with exponential rise and decay. The observed duration of the flare in this case is on the order of $d/c$. An example is shown in the {\it upper left} panel in fig. 2. Intermediate length outbursts can give rise to light curves that exhibit a fast rise, a plateau and a steep decline, as illustrated in the {\it upper right} and {\it bottom left} panels. Fig. 3 displays light curves at different observing frequencies (10, $10^3$, $10^5$, and $10^6$ GHz, with logarithmic energy intervals), produced for $d/r_o=10^{-2}$, and different values of $r_o$. As seen, the low frequency flux is delayed, owing to self absorption at the early stages of the outburst. Furthermore, the low frequency emission is strongly depressed when the radius of front formation, $r_o$, is sufficiently small. In the example depicted in fig. 3 there is essentially no radio emission for $r_o$ smaller than about 10, and very little emission even at millimeter wavelengths for $r_o<1$. The suppression of the flux at $10^6$ GHz seen for $r_o=10^2$ in the {\it upper left} panel in fig. 3, is due to our choice of $E_o$ and $B_o$ in eqs. (\ref{eq:Emx}) and (\ref{eq:B}). The total gamma-ray flux is emitted from roughly the same radii as the optically thin synchrotron flux and, therefore, correlations between the gamma-ray and optical/UV emission are expected. The precise relation, however, depends upon the combination of parameters through the radiative feedback and the evolution of the pair cascade, but the time lags should not exceed the flare duration. We emphasize that the gamma-ray flare can either lead or precede the optical/UV emission. As discussed in Paper I, the gamma-ray flux above the initial gamma-spheric radius should also be delayed and, for small enough $r_o$, will be strongly attenuated by pair production on external photons. Hence, we anticipate some correlated activity of the low-frequency and the hard gamma-ray emission (see further discussion below). The dependence of flare's properties on $p$ (defined in eq. [\ref{eq:B}]) is shown in fig. 4, for infinite length outbursts. As expected, increasing $p$ results in a substantially steeper decay of the emitted flux. As illustrated by this example, roughly symmetric flares can be produced essentially for a steep enough magnetic field profile, even in the case of thick slabs. The suppression of the low frequency flux at larger values of $p$, seen in the figure, is due to the steeper decline of the ratio of synchrotron and Compton cooling rates with time (see eq. [\ref {eq:cool}]). In the case of short length outbursts, the light curve is essentially insensitive to the radial profile of the magnetic field. Finally, we consider a situation whereby the maximum injection energy evolves with time. We then anticipate the development of the synchrotron flux at frequencies above the initial cutoff frequency, $\nu_{max}=(3eB/4\pi m_ec)E_o$ (cf. eq. [\ref{eq:Emx}]), to be governed by the evolution of $E_{emax}(t)$, whereas at lower frequencies the light curves should be highly insensitive to $E_{emax}(t)$, reflecting mainly the evolution of the synchrotron opacity, as in the previous example. This is illustrated in fig. 5, where light curves computed for infinite length outbursts and $b=1$ in eq. (\ref{eq:Emx}) are plotted. As seen, the onset of the flare at high frequencies becomes more delayed and the maximum flux becomes smaller for lower cutoff energies, $E_o$. \section{Summary and Conclusions} In this paper we considered the time dependent synchrotron emission produced by radiative fronts propagating in a magnetically dominated jet, and examined the relation between the synchrotron and ERC emission. Using a numerical model developed earlier, we analyzed the dependence of the variability pattern on the model parameters, and demonstrated how a change in physical parameters can lead to considerably different temporal behavior. We now summarize our results and conclusions. Two important parameters determine, in addition to the radial profile of the jet's magnetic field, the shape of the flare and the spectral evolution of the broad-band emission; the thickness of expelled fluid slab, and the radius at which the front is created. Ejection of sufficiently thin slabs leads to the production of roughly symmetric flares with exponential rise and decay, as occasionally seen in blazars (e.g., Massaro et al. 1996; Urry et al. 1997). The duration of the optically thin flare in this case should be on the order of the light crossing time of the ejected fluid slab, which can be as short as $r_{g}/c$, where $r_{g}$ is the gravitational radius of the putative black hole (cf. Paper I). The low frequency emission is quite generally delayed due to optical depth effects. Such delays provide an important diagnostic of radiatively efficient, inhomogeneous models in general, and are in agreement with the delays (of weeks to months) between high-energy and radio outbursts often observed (Reich et al. 1993; Wehrle et a.l, 1998), and the ejection of a superluminal blob following a gamma-ray flare (Zhang et al. 1994; Wehrle et al. 1996; Otterbein et al. 1998). The much shorter lags (hours to days) between the peaks of the gamma-ray, UV and optical fluxes, as seen on several occasions (e.g., Edelson et al. 1995; Urry et al. 1997; Wehrle et al., 1998), are also a natural consequence of this model. The precise relation between the optically thin emission in different bands depends on the conditions in the source through the radiative feedback and the evolution of pair cascades inside the front. If, in addition to being thin, the front is formed at a sufficiently small radius, then the low-frequency synchrotron flux will be strongly suppressed by virtue of self absorption, and the gamma-ray flux at energies above the corresponding gamma-spheric energy, will be severely attenuated by pair production on external photons ahead of the front. This implies that i) during such episodes a source may exhibit gamma-ray and UV/optical outbursts followed by little or no activity at low (typically radio-to-submillimeter) frequencies, and ii) in a single object the cutoff energy of the gamma-ray spectrum should be smaller for shorter outbursts, and should be correlated with the lack of activity (or with the amplitude of variations) at long wavelengths. Successful detection of such correlations in a source, particularly ii), will provide a strong support to this model, as such a behavior is not expected in SSC models or other types of inhomogeneous models and is, therefore, distinctive. For a reasonable choice of parameters we estimate the cutoff energy of the gamma-ray spectrum in the powerful blazars (which are likely to be ERC dominated) to lie in the range between a few GeV to a few hundreds GeV. It may be possible to observe it with the next generation gamma-ray telescope or, perhaps, with upcoming ground based experiments. Intermediate length outbursts can produce synchrotron light curves that exhibit a relatively rapid rise, a plateau and a sharp decay. We argue that the light curves reported by the PKS 2155, 1991 campaign (Urry et al. 1993; Courvoisier et al. 1995) might have been produced by such an event. A rise on a time scale of order a few days followed by a period of about 20 days during which the average flux remained at maximum level is evident (although the flux fluctuated around the maximum level during this period). Unfortunately, there is no data at later times. The faster, small amplitude oscilations may be caused by subsequent formation of smaller fronts, by instabilities, or by inhomogeneities in the upstream flow (which would lead essentially to bifurcation of the front). Furthermore, there is indication (Edelson et al. 1995) that the variation of the radio emission is delayed and has a smaller amplitude, consistent with this model. It is also tempting to relate the difference in temporal behavior seen by the 1991 and 1994 campaigns to the difference in conditions, particularly the length of the outburst, during those two episodes (cf. Georganopoulos \& Marscher 1997). When the thickness of expelled fluid slab largely exceeds the formation radius, which happens when the injection time of the jet by the central engine is much longer than the fluid acceleration time, the resultant light curve will exhibit an exponential rise and a power law decline. If the radial decrement of the jet's magnetic field is not too steep ($p<2$ in ep. [\ref{eq:B}]) then the flare will appear asymmetric, with a decay much longer than the rise when viewing the source at small enough angles. At larger viewing angles the shape of the flare may be altered by orientation effects (Eldar \& Levinson, in preparation). Such asymmetric flares are atypical to blazars, but are characteristic to GRB afterglows. In situations where the maximum injection energy increases with time, the onset of outbursts at frequencies above the initial upper cutoff becomes delayed, with longer time lags and smaller peak fluxes at higher frequencies, in contrast to the tendency found in the case of time independent $E_{emax}$. The gamma-ray light curve should not be affected significantly though (Levinson 1996), in contrast to the evolution predicted by SSC type models. Such events can lead to delayed, lower amplitude variations at optical/UV wavelengths, or even to the lack of apparent variations at these frequencies. Since the evolution of $E_{emax}$ is dictated by the rate of electron acceleration in the front, careful examination of the correlations at short wavelengths can provide important information regarding the acceleration process. Finally, we note that the inclusion of the SSC process may alter our results somewhat, particularly for parameters typical to fainter sources. The study of SSC flares is left for future work. I thank Avigdor Eldar for useful discussions, and the anonymous referee for constructive criticism. This research was supported by a grant from the Israeli Science Foundation and by Alon Fellowship.
\section*{Acknowledgements} I would like to thank T.~Jacobson, L.~J.~Mason, J.~M.~Nester, D.~C.~Robinson, R.~W.~Tucker and R.~M.~Wald for helpful comments; the Mathematics Department, King's College London, the Physics Department, National Central University, Lancaster University, and the Mathematical Institute at Oxford for hospitality while some of the work was carried out. The research was supported in part by NSC88-2112-M-008-018 (Taiwan) and EPSRC (UK).
\section{Introduction} The purpose of this paper is to point out a link between two apparently remote concepts: renormalization and Runge-Kutta methods. Renormalization enables us to remove infinities from quantum field theory. Recently, Kreimer discovered a Hopf algebra of rooted trees that brings order and beauty in the intricate combinatorics of renormalization \cite{Kreimer98}. He established formulas that automate the subtraction of infinities to all orders of the perturbation expansion, and proved the effectiveness of his method for the practical computation of renormalized quantities in joint works with Broadhurst \cite{Broadhurst} and Delbourgo \cite{Delbourgo}. Moreover, his approach shines new light on the problem of overlapping divergences \cite{KreimerOD,Krajewski} and on the mechanics of the renormalization group \cite{Kreimer}. Furthermore, Connes and Kreimer revealed a deep connection between the algebra of rooted trees (ART) and a Hopf algebra of diffeomorphisms \cite{Connes}. On the other hand, Runge published in 1895 \cite{Runge} an efficient algorithm to compute the solution of ordinary differential equations. For an equation of the type $dy/ds=f(y(s))$, he defines recursively $k_1=f(y_n)$, $k_2=f(y_n+h k_1/2)$, $y_{n+1}=y_n+hk_2$. His algorithm was improved in 1901 by Kutta, and became known as the Runge-Kutta method. It is now one of the most widely used numerical methods. In 1972, Butcher published an extraordinary article where he analyzed general Runge-Kutta methods on the basis of the ART. He showed that the Runge-Kutta methods form a group\footnote{Hairer and Wanner called it the Butcher group \cite{Hairer74}.} and found explicit expressions for the inverse of a method or the product of two methods. He also defined sums over trees that are now called B-series in honour of Butcher. Altough the Hopf algebra structure of ART is implicit all along his paper, Butcher did not mention it\footnote{The opposite of the present Hopf structure of ART was discussed in 1986 by D\"ur (\cite{Dur}, p.88-90), together with the corresponding Lie algebra (identical, up to a sign, with the one defined in \cite{Connes}).}. Important developments were made in 1974 by Hairer and Wanner \cite{Hairer74}. Since then, B-series are used routinely in the analysis of Runge-Kutta methods. Our main purpose is to show that the results and concepts established by Kreimer fit nicely into the Runge-Kutta language, and that the tools developed by Butcher have a range of application much wider than the numerical analysis of ordinary differential equations. The present expository paper will be reasonably self-contained. After an introduction to rooted trees, the genetic relation between ART and differentials is presented. Then Butcher's approach to Runge-Kutta methods is sketched. Several B-series are calculated and the connection with the Hopf structure of ART is exhibited. The application of Runge-Kutta methods to renormalization is exposed using a toy model which is solved non perturbatively. Finally, the solution of non-linear partial differential equations is written as a formal B-series. \section{The rooted trees} A rooted tree is a graph with a designated vertex called a root such that there is a unique path from the root to any other vertex in the tree \cite{Tucker}. Several examples of rooted trees are given in the appendix, where the root is the black point and the other vertices are white points (the root is at the top of the tree). The length of the unique path from a vertex $v$ to the root is called the level number of vertex $v$. The root has level number 0. For any vertex $v$ (except the root), the father of $v$ is the unique vertex $v'$ with an edge common with $v$ and a smaller level number. Conversely, $v$ is a son of $v'$. A vertex with no sons is a leaf. Rooted trees are sometimes called pointed trees or arborescences. The tree with one vertex is $\parbox{3mm}{$\bullet$}$, the ``tree'' with zero vertex is designated by $1$. \subsection{Operations and functions on trees} An important operation is the merging of trees. If $t_1$,\dots,$t_k$ are trees, $t=B^+(t_1,t_2,\dots,t_k)$ is defined as the tree obtained by creating a new vertex $r$ and by joining the roots of $t_1$,\dots,$t_k$ to $r$, which becomes the root of $t$. This operation is also denoted by $t=[t_1,t_2,\dots,t_k]$, but we avoid this notation because of the possible confusion with commutators. In \cite{Connes}, Connes and Kreimer defined a natural growth operator $N$ on trees: $N(t)$ is the set of $|t|$ trees $t_i$, where each $t_i$ is a tree with $|t|+1$ vertices obtained by attaching an additional leaf to a vertex of $t$. For example $N(1)=\parbox{3mm}{$\bullet$}$, \begin{eqnarray*} N(\parbox{3mm}{$\bullet$})&=&\tdeux,\quad N(\tdeux)=\ttroisun+\ttroisdeux\quad N(\ttroisdeux)=\tquatrequatre+2\,\tquatretrois. \end{eqnarray*} Some trees may appear with multiplicity. A number of functions on rooted trees have been defined independently by several authors: $|t|$ designates the number of vertices of a tree $t$ (alternative notation is $r(t)$, $\rho(t)$ and $\# t$). Clearly, $|B^+(t_1,t_2,\dots,t_k)|=|t_1|+|t_2|+\cdots+|t_k|+1$. The tree factorial $t!$ is defined recursively as \begin{eqnarray*} \parbox{3mm}{$\bullet$} !&=&1\\ B^+(t_1,t_2,\dots,t_k)!&=&|B^+(t_1,t_2,\dots,t_k)|\, t_1! t_2! \cdots t_k!. \end{eqnarray*} (an alternative notation is $\gamma(t)$). The notation $t!$ is taken from Kreimer \cite{Kreimer} because $t!$ generalizes the factorial of a number. Besides $t!$ has also similarities with the product of hooklengthes of a Young diagram in the representation theory of the symmetric group \cite{Fomin}. A few examples may be useful \begin{eqnarray*} \tdeux !&=&2,\quad \ttroisun !=6,\quad \tquatredeux != 12,\quad \tquatreun != 24, \quad\tquatrequatre !=4. \end{eqnarray*} \subsection{On $CM(t)$} $CM(t)$ was defined in \cite{Kreimer} as the number of times tree $t$ appears in $N^n(1)$ where $n=|t|$ is the number of vertices of $t$. In the literature (\cite{Butcher63}, \cite{Butcher}, p.92, \cite{Hairer}, p.147), $CM(t)$ is written as $\alpha(t)$ and considered as the number of ``heap-ordered trees'' with shape $t$, where a heap-ordered tree with shape $t$ is a labelling of each vertex of $t$ (i.e. a bijection between the vertices and the set of numbers $0,1,\dots |t|-1$) such that the labels decrease along the path going from any vertex to the root. This is called a monotonic labelling in \cite{Hairer}, p.147. \begin{eqnarray*} \parbox{10mm}{ \begin{fmfchar*}(10,10) \fmftop{i1} \fmfforce{(0,0)}{o2} \fmfforce{(0,0.6h)}{o1} \fmfright{o3} \fmf{vanilla}{i1,o1} \fmf{vanilla}{o1,o2} \fmf{vanilla}{i1,o3} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$0$},label.angle=0,label.dist=0,decor.size=90}{i1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$1$},label.angle=0,label.dist=0,decor.size=90}{o1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$2$},label.angle=0,label.dist=0,decor.size=90}{o2} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$3$},label.angle=0,label.dist=0,decor.size=90}{o3} \end{fmfchar*} } \quad\quad \parbox{10mm}{ \begin{fmfchar*}(10,10) \fmftop{i1} \fmfforce{(0,0)}{o2} \fmfforce{(0,0.6h)}{o1} \fmfright{o3} \fmf{vanilla}{i1,o1} \fmf{vanilla}{o1,o2} \fmf{vanilla}{i1,o3} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$0$},label.angle=0,label.dist=0,decor.size=90}{i1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$2$},label.angle=0,label.dist=0,decor.size=90}{o1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$3$},label.angle=0,label.dist=0,decor.size=90}{o2} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$1$},label.angle=0,label.dist=0,decor.size=90}{o3} \end{fmfchar*} } \quad\quad \parbox{10mm}{ \begin{fmfchar*}(10,10) \fmftop{i1} \fmfforce{(0,0)}{o2} \fmfforce{(0,0.6h)}{o1} \fmfright{o3} \fmf{vanilla}{i1,o1} \fmf{vanilla}{o1,o2} \fmf{vanilla}{i1,o3} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$0$},label.angle=0,label.dist=0,decor.size=90}{i1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$1$},label.angle=0,label.dist=0,decor.size=90}{o1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$3$},label.angle=0,label.dist=0,decor.size=90}{o2} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$2$},label.angle=0,label.dist=0,decor.size=90}{o3} \end{fmfchar*} } \end{eqnarray*} There are $(n-1)!$ heap-ordered trees with $n$ vertices. This can be seen by a recursive argument. Take a heap-ordered tree $t$ with $n$ vertices, and make $n$ labeled trees by adding a new vertex with label $n$ to each vertex of $t$. Then all the created trees are heap-ordered (because the added vertex is a leaf and all other labels are smaller than $n$). Furthermore, all the heap-ordered trees created by this process from the set of heap-ordered trees with $n$ vertices are different. Therefore, there are at least $n!$ heap-ordered trees with $n+1$ vertices. On the other hand, in each heap-ordered tree $t$ with $n+1$ vertices, the vertex labeled $n$ is a leaf, therefore $t$ can be created from a heap-ordered tree with $n$ vertices by adding this leaf with label $n$. So there are exactly $n!$ heap-ordered trees with $n+1$ vertices. We shall give a non combinatiorial proof of this fact in the sequel. Since $N(t)$ is defined by the addition of a leaf to all the vertices of $t$, $\alpha(t)$ is the number of heap-ordered trees with shape $t$. This number has been calculated in \cite{Butcher63} (see also. \cite{Butcher}, p.92): \begin{eqnarray*} \alpha(t)=\frac{|t|!}{t!S_t}, \end{eqnarray*} where $S_t$ is the symmetry factor of $t$, defined in \cite{Broadhurst,Kreimer} and in \cite{Butcher} where it is denoted by $\sigma(t)$. Note that there is a simple correspondence between the permutations of $n-1$ numbers and the heap-ordered trees. Let $(p_1,\dots,p_{n-1})$ be a permutation of $(1,\dots,n-1)$, then \begin{itemize} \item $p_1$ is a subroot, labeled $p_1$ \item for $i$=2 to $n-1$ \begin{itemize} \item if all $p_j$ for $1\le j\le i$ are such that $p_j > p_i$, then $p_i$ is a subroot, labeled $p_i$ \item otherwise, let $p_j$ be first number such that $p_j < p_i$, in the series $p_{i-1},p_{i-2},\dots,p_1$, then the $i$-th vertex, labeled $p_i$, is linked to $p_j$ by a line \end{itemize} \item when all $(p_1,\dots,p_{n-1})$ have been processed, all subroots are linked to a common root, labeled $0$ \end{itemize} On the other hand, starting from a heap-ordered tree $t$, $t$ is arranged so that the set of all vertices with a given level number are ordered with labels increasing from right to left. Then the permutation is built by gathering the labels through a depth-first search (backtracking) of the tree from left to right. For instance, the permutation corresponding to the three labeled trees of the above example are (312), (231) and (213). Finally, we use the term algebra of rooted trees and not Hopf algebra of rooted trees because, thanks to the work of Butcher, the Hopf structure is only one aspect of the ART. \section{Differentials and rooted trees} Assume that we want to solve the equation $(d/ds)x(s)=F[x(s)]$, $x(s_0)=x_0$, where $s$ is a real, $x$ is in $\mathbb{R}^N$ and $F$ is a smooth function from $\mathbb{R}^n$ to $\mathbb{R}^N$, with components $f^i(x)$. This is the equation of the flow of a vector field. \subsection{Calculation of the $n$-th derivative} Let us write the derivatives of the $i$-th component of $x(s)$ with respect to $s$: \begin{eqnarray*} \frac{d^2x^i(s)}{ds^2}&=&\frac{d}{ds} f^i[x(s)] = \sum_j\frac{\partial f^i}{\partial x_j}[x(s)] \frac{dx^j}{ds} = \sum_j\frac{\partial f^i}{\partial x_j}[x(s)] f^j[x(s)] \end{eqnarray*} \begin{eqnarray*} \frac{d^3x^i(s)}{ds^3}&=&\frac{d}{ds} \left(\sum_j\frac{\partial f^i}{\partial x_j}[x(s)] f^j[x(s)]\right)\\ &=& \sum_{jk}\frac{\partial^2 f^i}{\partial x_j\partial x_k}[x(s)] f^j[x(s)] f^k[x(s)] + \sum_{jk}\frac{\partial f^i}{\partial x_j}[x(s)] \frac{\partial f^j}{\partial x_k}[x(s)] f^k[x(s)]. \end{eqnarray*} A simplified notation is now required. Let \begin{eqnarray*} f^i &=& f^i[x(s)] \\ f^i_{j_1j_2\cdots j_k} &=& \frac{\partial^k f^i}{\partial x_{j_1}\cdots\partial x_{j_k}}[x(s)], \end{eqnarray*} so that \begin{eqnarray*} \frac{dx^i(s)}{ds}&=& f^i\quad \frac{d^2x^i(s)}{ds^2}= f^i_j f^j\quad \frac{d^3x^i(s)}{ds^3}= f^i_{jk} f^j f^k + f^i_j f^j_k f^k, \end{eqnarray*} where summation over indices appearing in lower and upper positions is implicitly assumed. With this notation, we can write the next term as \begin{eqnarray*} \frac{d^4x^i(s)}{ds^4}&=& f^i_j f^j_k f^k_l f^l + f^i_j f^j_{kl} f^k f^l + 3 f^i_{jk} f^j_l f^k f^l + f^i_{jkl} f^j f^k f^l \\ &=& \parbox{10mm}{ \begin{fmfchar*}(10,12) \fmftop{i1} \fmfbottom{o1} \fmf{vanilla}{i1,v1} \fmf{vanilla}{v1,v2} \fmf{vanilla}{v2,o1} \fmfv{decor.shape=circle,decor.filled=1,% label=\mbox{\small$i$},label.angle=0,label.dist=20,decor.size=30}{i1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$j$},label.angle=0,label.dist=20,decor.size=30}{v1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$k$},label.angle=0,label.dist=20,decor.size=30}{v2} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$l$},label.angle=0,label.dist=20,decor.size=30}{o1} \end{fmfchar*} } \quad \quad \quad \parbox{6mm}{ \begin{fmfchar*}(6,10) \fmftop{i1} \fmfbottom{o1,o2} \fmfforce{(0.5w,0.5h)}{v1} \fmf{vanilla}{i1,v1} \fmf{vanilla}{v1,o1} \fmf{vanilla}{v1,o2} \fmfv{decor.shape=circle,decor.filled=1,% label=\mbox{\small$i$},label.angle=0,label.dist=20,decor.size=30}{i1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$j$},decor.size=30}{v1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$k$},decor.size=30}{o1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$l$},decor.size=30}{o2} \end{fmfchar*} } \quad \quad \quad \quad \parbox{6mm}{ \begin{fmfchar*}(6,10) \fmftop{i1} \fmfforce{(0,0)}{o2} \fmfforce{(0,0.5h)}{o1} \fmfright{o3} \fmf{vanilla}{i1,o1} \fmf{vanilla}{o1,o2} \fmf{vanilla}{i1,o3} \fmfv{decor.shape=circle,decor.filled=1,% label=\mbox{\small$i$},label.angle=0,label.dist=20,decor.size=30}{i1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$j$},decor.size=30}{o1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$l$},decor.size=30}{o2} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$k$},decor.size=30}{o3} \end{fmfchar*} } \quad \quad \quad \quad \parbox{8mm}{ \begin{fmfchar*}(8,6) \fmftop{i1} \fmfstraight \fmfbottom{o1,o2,o3} \fmf{vanilla}{i1,o1} \fmf{vanilla}{i1,o2} \fmf{vanilla}{i1,o3} \fmfv{decor.shape=circle,decor.filled=1,% label=\mbox{\small$i$},decor.size=30}{i1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$j$},decor.size=30}{o1} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$k$},decor.size=30}{o2} \fmfv{decor.shape=circle,decor.filled=0,% label=\mbox{\small$l$},decor.size=30}{o3} \end{fmfchar*} } \end{eqnarray*} This relation between differentials and rooted tree was established by Arthur Cayley in 1857 \cite{Cayley}. With this notation, there is a one-to-one relation between a rooted tree with $n$ vertices and a term of $d^n x(s)/ds^n$ \subsection{Elementary differentials} A little bit more formally, we can follow Butcher (\cite{Butcher}, p.154.) and call ``elementary differentials'' the $\delta_t$ defined recursively for each rooted tree $t$ by: \begin{eqnarray} \delta^i_\bullet &=& f^i \nonumber\\ \delta^i_t &=& f^i_{j_1j_2\cdots j_k} \delta^{j_1}_{t_1} \delta^{j_2}_{t_2} \cdots \delta^{j_k}_{t_k} \quad\mathrm{when}\quad t=B^+(t_1,t_2,\cdots,t_k). \label{defdeltat} \end{eqnarray} Using this correspondence between rooted trees and differential expressions, we establish the identity: \begin{eqnarray*} N\delta_{t}\equiv\frac{d\delta_t}{ds}=\delta_{N(t)}, \end{eqnarray*} where $N(t)$ is the natural growth operator of rooted trees defined in Ref.\cite{Connes}. So that the solution of the flow equation is \begin{eqnarray} x(s)&=&x_0+\int_{s_0}^s ds' \exp[s'N]\delta_\bullet\nonumber\\ &=&x_0+\sum_t \frac{(s-s_0)^{|t|}}{|t|!} \alpha(t) \delta_t(s_0), \label{Butcherflow} \end{eqnarray} where $|t|$ is the number of vertices of $t$, and $\alpha(t)$ is called $CM(t)$ in \cite{Kreimer}. \section{Runge-Kutta methods} We shall see that sum over trees appear quite naturally with differential equations. So, if one is given a function $\phi$ that assigns a value (e.g. a real, a complex, a vector) to each tree $t$, is there a function $f$ such that $\phi(t)=\delta_t$. Generally, the answer is no. Consider a function $\phi$ such that all components are equal (and denoted also by $\phi$): \begin{eqnarray*} \phi(\parbox{3mm}{$\bullet$})&=&1,\quad \phi(\tdeux)=a,\quad \phi(\ttroisun)=b, \end{eqnarray*} so that for any $i$, $f^i=1$, $f^i_j f^j=a$ and $f^i_j f^j_kf^k=b$. The first two equations give $\sum_jf^i_j=a$, so that the third gives $f^i_j f^j_kf^k=\sum_j f^i_j a=a^2$, and $\phi$ cannot be represented as elementary differentials (i.e. it cannot be the $\delta_t$) of a function $f$ if $b\not=a^2$. In fact, the number of functions reachable as elementary differentials is rather narrow. Given such a function $\phi$ over rooted trees, we extend it to a homomorphism of the algebra of rooted trees by linearity and $\phi(tt')=\phi(t)\phi(t')$ where the componentwise product was used on the right-hand side. If vector flows are not enough to span all possible $\phi$, what more general equation can do that? As we shall see now, the answer is the Runge-Kutta methods. \subsection{Butcher's approach to the Runge-Kutta methods} To solve a flow equation $dx(s)/ds=F[x(s)]$, some efficient numerical algorithms are known as Runge-Kutta methods. They are determined by a $m\times m$ matrix $a$ and an $m$-dimensional vector $b$, and at each step a vector $x_n$ is defined as a function of the previous value $x_{n-1}$ by: \begin{eqnarray*} X_i &=& x_{n-1}+h\sum_{j=1}^m a_{ij} F(X_j)\\ x_n &=& x_{n-1}+h\sum_{j=1}^m b_j F(X_j), \end{eqnarray*} where $i$ range from $1$ to $m$. If the matrix $a$ is such that $a_{ij}=0$ if $j\ge i$ then the method is called explicit (because each $X_i$ can be calculated explicitly), otherwise the method is implicit. In 1963, Butcher showed that the solution of the corresponding equations: \begin{eqnarray*} X_i(s) &=& x_0+(s-s_0)\sum_{j=1}^m a_{ij} F(X_j(s))\\ x(s) &=& x_0+(s-s_0)\sum_{j=1}^m b_j F(X_j(s)), \end{eqnarray*} is given by \begin{eqnarray} X_i(s)&=&x_0+\sum_t \frac{(s-s_0)^{|t|}}{|t|!} \alpha(t) t! \sum_{j=1}^m a_{ij} \phi_j(t) \delta_t(s_0) \nonumber\\ x(s)&=&x_0+\sum_t \frac{(s-s_0)^{|t|}}{|t|!} \alpha(t) t! \phi(t) \delta_t(s_0). \label{Bseries} \end{eqnarray} These series over trees are called B-series in the numerical analysis literature, in honour of John Butcher (\cite{Hairer}, p.264). The homomorphism $\phi$ is defined recursively as a function of $a$ and $b$, for $i=1,\dots,m$: \begin{eqnarray*} \phi_i(\parbox{3mm}{$\bullet$})&=&1\\ \phi_i(B^+(t_1\cdots t_k))&=&\sum_{j_1,\dots,j_k} a_{ij_1}\dots a_{ij_k}\phi_{j_1}(t_1)\dots\phi_{j_k}(t_k)\\ \phi(t)&=&\sum_{i=1}^m b_i \phi_i(t). \end{eqnarray*} Comparing Eqs.(\ref{Butcherflow}) and (\ref{Bseries}) it is clear that the Runge-Kutta approximates the solution of the original flow equation up to order $n$ if $\phi(t)=1/t!$ for all trees with up to $n$ vertices. In 1972 \cite{Butcher72}, Butcher made further progress. Firstly he showed that Runge-Kutta methods are ``dense'' in the space of rooted tree homomorphisms. More precisely, he showed that given any finite set of trees $T_0$ and any function $\theta$ from $T_0$ to $\mathbb{R}$, then there is a Runge-Kutta method (i.e. a matrix $a$ and a vector $b$) such that the corresponding $\phi$ agrees with $\theta$ on $T_0$ (see also \cite{Butcher} p.167). \subsection{Further developments} Furthermore, Butcher proved that the combinatorics he used to study Runge-Kutta methods in 1963 \cite{Butcher63} was hiding an algebra. If ($a$,$b$) and ($a'$,$b'$) are two Runge-Kutta methods, with the corresponding homomorphisms $\phi$ and $\phi'$, then the product homomorphism is defined (in Hopf algebra terms) by \begin{eqnarray*} (\phi\star\phi')(t)&=&m[(\phi\otimes\phi')\Delta(t)]. \end{eqnarray*} Butcher proved that the $\phi$ derived from Runge-Kutta methods form a group. Again, this is nicely interpreted within the Hopf structure of the ART. For instance, the inverse of the element $\phi$ is simply defined by $\phi^{-1}(t)=\phi[S(t)]$, where $S$ is the antipode. This concept of inverse is quite important in practice since it is involved in the concept of self-adjoint Runge-Kutta methods, which have long-term stability in time-reversal symmetric problems (\cite{Hairer}, p.219). The adjoint is defined within our approach by $\phi^*(t)=(-1)^{|t|}\phi[S(t)]$. On the other hand, Butcher found an explicit expression for all the Hopf operations of the ART. Given the method ($a$,$b$) for $\phi$, he expressed the method ($a'$,$b'$) for $\phi\circ S$ ($S$ is the antipode) in (simple) terms of ($a$,$b$). Moreover, (\cite{Butcher}, p.312 et sq.), if ($a$,$b$) and ($a'$,$b'$) are two Runge-Kutta methods (with dimensions $m$ and $m'$, respectively), corresponding to $\phi$ and $\phi'$, the method ($a"$,$b"$) corresponding to the convolution product $(\phi\star\phi')$ is \begin{eqnarray*} a''_{ij}&=&a_{ij}\mathrm{\quad if\quad} 1\le i\le m \mathrm{\quad and\quad} 1\le j\le m,\\ a''_{ij}&=&a'_{ij}\mathrm{\quad if\quad} m+1\le i\le m+m' \mathrm{\quad and\quad} m+1\le j\le m+m',\\ a''_{ij}&=&b_{j}\mathrm{\quad if\quad} m+1\le i\le m+m' \mathrm{\quad and\quad} 1\le j\le m,\\ a''_{ij}&=&0 \mathrm{\quad if\quad} 1\le i\le m \mathrm{\quad and\quad} m+1\le j\le m+m',\\ b''_i &=& b_i \mathrm{\quad if\quad} 1\le i\le m,\\ b''_i &=& b'_i \mathrm{\quad if\quad} m+1\le i\le m+m'. \end{eqnarray*} In 1974, Hairer and Wanner (\cite{Hairer}, p.267) built upon the work of Butcher and proved the following important result: if we denote \begin{eqnarray} B(\phi,F)&=&1+\sum_t \frac{(s-s_0)^{|t|}}{|t|!} \alpha(t) t! \phi(t) \delta_t(s_0) \label{Hairer1} \end{eqnarray} then \begin{eqnarray} B(\phi',B(\phi,F))&=&B(\phi\star\phi',F), \label{Hairer2} \end{eqnarray} where $B(\phi',B(\phi,F))$ is the same as Eq.(\ref{Hairer1}), with $\phi(t)$ replaced by $\phi'(t)$ and $\delta_t$ replaced by $\delta'(t)$ (i.e. $\delta'(t)$ is calculate as $\delta_t$, but with the function $B(\phi,F)(s)$ instead of the function $F(x(s))$). In other words, the group of homomorphisms acts on the right on the functions $F$. \section{The continuous limit} In his seminal article \cite{Butcher72}, Butcher did not restrict his treatment to finite sets of indices. It is possible to consider the continuous limit of Runge-Kutta methods. A possible form of it is an integral equation, which we write artitrarily between 0 and 1: \begin{eqnarray*} X_u(s) &=& x_0+(s-s_0)\int_0^1 dv a(u,v) F(X_v(s))\\ x(s) &=& x_0+(s-s_0)\int_0^1 b(u) du F(X_u(s)), \end{eqnarray*} the solution of which are \begin{eqnarray*} X_u(s)&=&x_0+\sum_t \frac{(s-s_0)^{|t|}}{|t|!} \alpha(t) t! \int_0^1 dv a(u,v) \phi_v(t) \delta_t(s_0)\\ x(s)&=&x_0+\sum_t \frac{(s-s_0)^{|t|}}{|t|!} \alpha(t) t! \phi(t) \delta_t(s_0). \end{eqnarray*} The homomorphism $\phi$ is defined recursively as a function of $a$ and $b$: \begin{eqnarray*} \phi_u(\parbox{3mm}{$\bullet$})&=&1\\ \phi_u(B^+(t_1\cdots t_k))&=&\int_0^1 du_1 a(u,u_1) \phi_{u_1}(t_1) \dots\int_0^1 du_k a(u,u_k) \phi_{u_k}(t_k)\\ \phi(t)&=&\int_0^1 du b(u) \phi_u(t). \end{eqnarray*} Continuous RK-methods do not seem to have been much used, except for an example in Butcher's book (\cite{Butcher} p.325). \subsection{Butcher's example \label{Butcherexample}} It will be useful in the following to have the results of a modified version of Butcher's example. So, we consider: \begin{eqnarray} X_u(s)&=&x_0+(s-s_0)\int_0^u F[X_v(s)] dv \label{Picard}\\ x(s)&=&x_0+(s-s_0)\int_0^1 F[X_u(s)] du\nonumber, \end{eqnarray} which corresponds to $a(u,v)=1_{[0,u]}(v)$, $b(u)=1$. This Runge-Kutta method will be used again in the sequel, and will be referred to as the ``simple integral method''. If we take the derivative of Eq.(\ref{Picard}) with respect to $u$ we obtain \begin{eqnarray*} \frac{d}{du} X_u(s)=(s-s_0) F[X_u(s)], \end{eqnarray*} so $X_u(s)=y(s_0+(s-s_0)u)$, where $y(s)$ is the solution \begin{eqnarray*} y(s)=x_0+\int_{s_0}^s F[y(s')] ds'. \end{eqnarray*} Moreover \begin{eqnarray*} x(s)&=&x_0+(s-s_0)\int_0^1 F[X_u(s)] du\\ &=&x_0+(s-s_0)\int_0^1 F[y(s_0+(s-s_0)u)] du\\ &=&x_0+\int_{s_0}^s F[y(s')] ds' = y(s). \end{eqnarray*} The corresponding homomorphism $\phi(t)$ is defined by \begin{eqnarray*} \phi_u(\parbox{3mm}{$\bullet$})&=&1\\ \phi_u(B^+(t_1\cdots t_k))&=&\int_0^u du_1 \phi_{u_1}(t_1) \dots\int_0^u du_k \phi_{u_k}(t_k)\\ \phi(t)&=&\int_0^1 du \phi_u(t). \end{eqnarray*} Using the facts that $|B^+(t_1\cdots t_k)|=(|t_1|+\cdots+|t_k|+1)$ and $B^+(t_1\cdots t_k)!=(|t_1|+\cdots+|t_k|+1)t_1!\dots t_k!$ it is proved that the solutions of these equations are \begin{eqnarray*} \phi_u(t)&=&\frac{|t|u^{|t|-1}}{t!}\\ \phi(t)&=&\frac{1}{t!}. \end{eqnarray*} If we introduce $\phi(t)=1/t!$ into Eq.(\ref{Bseries}) we obtain Eq.(\ref{Butcherflow}). So we confirm that the solution of the equation \begin{eqnarray*} x(s)=x_0+\int_{s_0}^s F[x(s')] ds' \end{eqnarray*} is \begin{eqnarray*} x(s)&=&x_0+ \sum_t \frac{(s-s_0)^{|t|}}{|t|!} \alpha(t) \delta_t(s_0). \end{eqnarray*} \subsection{First applications} The above example can already bring some interesting applications. But we must start by giving a way to calculate $\delta_t(s_0)$ in a simple case. \subsubsection{Calculation of $\delta_t(s_0)$ \label{sectiondeltat}} To obtain specific results, we must choose a particular function $F$. The simplest choice is to take a vector function $F$, with components $f^i(x)=f(\sum_j x_j/N)$, where $N$ is the dimension of the vector space and $f$ has the series expansion \begin{eqnarray*} f(s)=\sum_{n=0}^\infty \frac{f^{(n)}(0) s^n}{n!}. \end{eqnarray*} From the definition of $\delta_t$ in Eq.(\ref{defdeltat}), one can show recursively that, for $i=1,\dots,N$, $\delta^i_t(0)$ is independent of $i$ (and will be denoted $\delta_t$) and \begin{eqnarray} \delta_\bullet &=& f(0) \nonumber\\ \delta_t &=& f^{(k)}(0) \delta_{t_1} \delta_{t_2} \cdots \delta_{t_k} \quad\mathrm{when}\quad t=B^+(t_1,t_2,\cdots,t_k). \label{defdeltatsimple} \end{eqnarray} In ref.\cite{Kreimer}, Kreimer defined a similar quantity, that he called $B_t$. Here $\delta_t$ and $B_t$ will be used as synonymous. The simplest case is $f(s)=\exp s$ and $s_0=0$, where $f^{(n)}(0)=1$ and $\delta_t=1$ for all trees $t$. \subsubsection{Weighted sum of rooted trees} If we take $f=\exp$, $s_0=0$ and $x_0=0$ in Butcher's example (see section \ref{Butcherexample}), we have to solve the equation \begin{eqnarray*} x(s)=\int_{0}^s \exp[x(s')] ds' \end{eqnarray*} which can be differentiated to give $x'(s)=\exp(x(s))$ with $x(0)=0$. This has the solution \begin{eqnarray*} x(s)=-\log(1-s)=\sum_{n=1}^\infty \frac{s^n}{n}. \end{eqnarray*} On the other hand, the corresponding homomorphism is $\phi(t)=1/t!$ and the B-series for this problem is \begin{eqnarray*} x(s)&=& \sum_t \frac{s^{|t|}}{|t|!} \alpha(t). \end{eqnarray*} Comparing the last two results, we find \begin{eqnarray*} \sum_{|t|=n} \alpha(t)=(n-1)! \end{eqnarray*} in other words, the number of heap-ordered trees with $n$ vertices is $(n-1)!$. \subsubsection{Derivative of inverse functions} We can try to extend the last example to an arbitrary function $f(x)$. The equation to solve becomes \begin{eqnarray*} x(s)=\int_{0}^s f[x(s')] ds', \end{eqnarray*} or $x'(s)=f(x(s))$ with $x(0)=0$. Let \begin{eqnarray*} S(x)=\int_{0}^x \frac{dy}{f(y)}, \end{eqnarray*} which gives us $s=S(x)$, or $x(s)=S^{-1}(s)$, where $S^{-1}$ is the inverse function of $S$. If $f=\exp$, $S(x)=1-\exp(-x)$ and we confirm that $x(s)=-\log(1-s)$. We can use this result to calculate the derivatives of a function $x(s)$, given as the inverse of a function $S(x)$. To do this, we define $f(x)=1/S'(x)$ and, using Eq.(\ref{Butcherflow}), we obtain \begin{eqnarray} x^{(n)}(0) &=&\sum_{|t|=n} \alpha(t) \delta_t, \end{eqnarray} where $\delta_t$ is calculated from $f(s)$ using Eq.(\ref{defdeltatsimple}) in section \ref{sectiondeltat}. This method can also be calculated to find the function $f$ satisfying given values for \begin{eqnarray*} a_n=\sum_{|t|=n} \alpha(t) \delta_t, \end{eqnarray*} where $\delta_t$ is calculated from $f$. For instance, if we want \begin{eqnarray*} \sum_{|t|=n} \alpha(t) \delta_t=n!, \end{eqnarray*} we must take $f(s)=1+s^2$. \subsubsection{Other sums over trees \label{sectionothersum}} We give now further examples of sums over trees, that will be used in the sequel. For instance, assume that we need to compute \begin{eqnarray*} S=\sum_{|t|=n} \frac{\alpha(t)}{t!}. \end{eqnarray*} This term comes in the Butcher series with $\phi(t)=1/(t!)^2$. Since this $\phi(t)$ is the square of the previous one, the corresponding Runge-Kutta method can be realized as the tensor product of two ``simple integral methods'' (see section \ref{Butcherexample}). In other words \begin{eqnarray*} a(u,u',v,v')&=&a(u,v) a(u',v')=1_{[0,u]}(v)1_{[0,u']}(v') \quad b(u,u')=b(u)b(u')=1. \end{eqnarray*} and the Runge-Kutta method is now \begin{eqnarray*} X_{uu'}(s)&=&x_0+(s-s_0)\int_0^u dv \int_0^{u'} dv'f[X_{vv'}(s)] \\ x(s)&=&x_0+(s-s_0)\int_0^1du \int_0^1du' f[X_{uu'}(s)]. \end{eqnarray*} The corresponding homomorphism $\phi(t)$ is given by \begin{eqnarray*} \phi_{uu'}(\parbox{3mm}{$\bullet$})&=&1\\ \phi_{uu'}(B^+(t_1\cdots t_k))&=&\int_0^u du_1 \int_0^{u'} du'_1 \phi_{u_1u'_1}(t_1) \dots\int_0^u du_k \int_0^{u'} du'_k \phi_{u_ku'_k}(t_k)\\ \phi(t)&=&\int_0^1 du \int_0^1 du'\phi_{uu'}(t). \end{eqnarray*} The solutions of these equations are \begin{eqnarray*} \phi_{uu'}(t)&=&\frac{|t|^2(uu')^{|t|-1}}{(t!)^2}\\ \phi(t)&=&\frac{1}{(t!)^2}, \end{eqnarray*} so that, from Eq.(\ref{Bseries}) \begin{eqnarray*} X_{uu'}(s)&=&x_0+\sum_t \frac{(s-s_0)^{|t|}}{|t|!} \frac{\alpha(t)(uu')^{|t|}}{t!} \delta_t(s_0). \end{eqnarray*} The conclusion is that $X_{uu'}(s)$ is in fact a function of $uu'$ and not of $u$ and $u'$. More precisely, we know from the general formula Eq.(\ref{Bseries}) that the B-series for the solution of $x(s)=x_0+(s-s_0)\int_0^1du \int_0^1du' f[X_{uu'}(s)]$ is \begin{eqnarray*} x(s)&=&x_0+ \sum_t \frac{(s-s_0)^{|t|}}{|t|!} \frac{\alpha(t)}{t!} \delta_t(s_0), \end{eqnarray*} so that $X_{uu'}(s)=x(s_0+(s-s_0)uu')$. If we use the successive changes of variables $w=uu'$, $v'=s_0+(s-s_0)w$ and $v=s_0+(s-s_0)u$ we find \begin{eqnarray*} x(s)&=&x_0+(s-s_0)\int_0^1du \int_0^1du' f[x(s_0+(s-s_0)uu')]\\ &=&x_0+(s-s_0)\int_0^1\frac{du}{u} \int_0^u dw f[x(s_0+(s-s_0)w)]\\ &=&x_0+\int_0^1\frac{du}{u} \int_{s_0}^{s_0+(s-s_0)u} dv' f[x(v')]\\ &=&x_0+\int_{s_0}^{s}\frac{dv}{v-s_0} \int_{s_0}^{v} dv' f[x(v')]. \end{eqnarray*} With the initial values $x_0=s_0=0$ this gives us \begin{eqnarray} x(s) &=&\int_{0}^{s}\frac{dv}{v} \int_{0}^{v} dv' f[x(v')], \label{doublefact} \end{eqnarray} or $sx''+x'=f(x)$ with $x(0)=0$ and $x'(0)=f(0)$. If we take again $f(x)=\exp(x)$ we find $sx''+x'=\exp(x)$ with $x(0)=0$ and $x'(0)=1$, so that \begin{eqnarray*} x(s)=-2\log(1-s/2)=\sum_{n=1}^\infty \frac{s^n}{n2^{n-1}}. \end{eqnarray*} Comparing this with the B-series \begin{eqnarray*} x(s)&=& \sum_t \frac{s^{|t|}}{|t|!} \frac{\alpha(t)}{t!} \end{eqnarray*} we obtain \begin{eqnarray*} S=\sum_{|t|=n} \frac{\alpha(t)}{t!}=\frac{(n-1)!}{2^{n-1}}, \end{eqnarray*} which is the result found by Kreimer in \cite{Kreimer} using combinatorial arguments. As a final example, we can consider the Runge-Kutta method $a(u,v)=1$, $b(u)=1$ which gives $\phi(t)=1$ for all trees $t$. The equation for $x(s)$ is now a fixed point problem $x(s)=s\exp(x(s))$, whose well-known solution is \begin{eqnarray*} x(s)&=& \sum_n \frac{s^n}{n!} n^{n-1}, \end{eqnarray*} so that \begin{eqnarray*} \sum_{|t|=n} \alpha(t) t!=n^{n-1}. \end{eqnarray*} These examples show that B-series can be used as generating series for sums over trees. \subsection{The antipode \label{sectionantipode}} The Hopf algebra structure of the ART entails an antipode $S$. If $\phi(t)$ is an homomorphism, the action of the antipode on $\phi$ can be written as $S(\phi)(t)=\phi(S(t))$. If the Runge-Kutta method for $\phi$ is $A_u$, $B$, then the Runge-Kutta method for $\phi^S=S(\phi)$ is $A^S_u=A_u-B$, $B^S=-B$. It is useful to see it working on simple cases: \begin{eqnarray*} \phi^S_u(\parbox{3mm}{$\bullet$})&=&1=\phi_u(\parbox{3mm}{$\bullet$})\\ \phi^S(\parbox{3mm}{$\bullet$})&=&B^S(\phi^S_u(\parbox{3mm}{$\bullet$}))=-B(\phi_u(\parbox{3mm}{$\bullet$}))=-\phi(\parbox{3mm}{$\bullet$})\\ \phi^S_u(\tdeux)&=&A^S_u(\phi^S_v(\parbox{3mm}{$\bullet$}))=A_u(\phi_v(\parbox{3mm}{$\bullet$}))-B(\phi_v(\parbox{3mm}{$\bullet$}))= \phi_u(\tdeux)-\phi(\parbox{3mm}{$\bullet$})\\ \phi^S(\tdeux)&=&-B(\phi^S_u(\tdeux))=-\phi(\tdeux)+B(1)\phi(\parbox{3mm}{$\bullet$})= -\phi(\tdeux)+\phi(\parbox{3mm}{$\bullet$})\phi(\parbox{3mm}{$\bullet$}) \end{eqnarray*} \subsection{The convolution \label{sectionconvolution}} The convolution of $\phi$ and $\phi'$ is defined as $\phi''(t)=(\phi\star\phi')(t)=m[(\phi\otimes\phi')\Delta(t)]$. Let $A_u$, $B$ and $A'_u$, $B'$ be the Runge-Kutta methods of, respectively, $\phi(t)$ and $\phi'(t)$. To be specific, we consider that $u$ varies from 0 to 1. Then the Runge-Kutta method for $\phi''$ is $A''_u$, $B''$, where $u$ varies from 0 to 2 and \begin{eqnarray*} A''_u(X_v)&=&A_u(X_v)\quad\mathrm{if}\quad 0\le u\le 1 \quad\mathrm{and}\quad 0\le v\le 1\\ A''_u(X_v)&=&0\quad\mathrm{if}\quad 0\le u\le 1 \quad\mathrm{and}\quad 1\le v\le 2\\ A''_u(X_v)&=&B(X_v) \quad\mathrm{if}\quad 1\le u\le 2 \quad\mathrm{and}\quad 0\le v\le 1\\ A''_u(X_v)&=&A'_{u-1}(X_{v-1}) \quad\mathrm{if}\quad 1\le u\le 2 \quad\mathrm{and}\quad 1\le v\le 2\\ B''(X_v)&=&B(X_v)\quad\mathrm{if}\quad 0\le v\le 1\\ B''(X_v)&=&B'(X_{v-1})\quad\mathrm{if}\quad 1\le v\le 2. \end{eqnarray*} Again, we show the formula in action: \begin{eqnarray*} \phi''_u(\parbox{3mm}{$\bullet$})&=&1\\ \phi''(\parbox{3mm}{$\bullet$})&=&B(1)+B'(1)=\phi(\parbox{3mm}{$\bullet$})+\phi'(\parbox{3mm}{$\bullet$})\\ \phi''_u(\tdeux)&=&A''_u(\phi''_v(\parbox{3mm}{$\bullet$}))=A_u(1) 1_{[0,1](u)} +(B(1)+A'_{u-1}(1))1_{[1,2](u)}\\ \phi''(\tdeux)&=&B(A_u(1))+B'((B(1)+A'_{u-1}(1)))= B(\phi_u(\parbox{3mm}{$\bullet$}))+B(1)B'(1)+B'(\phi'u(\parbox{3mm}{$\bullet$}))\\ &=& \phi(\tdeux)+\phi(\parbox{3mm}{$\bullet$})\phi'(\parbox{3mm}{$\bullet$})+\phi'(\tdeux). \end{eqnarray*} \section{Runge-Kutta methods for renormalization} In this section, we shall follow closely Kreimer's paper \cite{Kreimer} and define, for each operation on homomorphisms, a corresponding transformation of the Runge-Kutta methods. Instead of attempting a general theory, we consider a specific example in detail. \subsection{Runge-Kutta method for bare quantities} We consider that a given bare physical quantity can be calculated as a sum over trees, and that the corresponding Runge-Kutta method has been found as a pair of linear operators $A_u$ and $B$. The usual combinatorial proof show that the solution of the equations (we take $s_0=0$) \begin{eqnarray*} X_u(s) &=& x_0+s A_u[f(X_v(s))]\\ x(s) &=& x_0+s B[f(X_u(s))], \end{eqnarray*} is \begin{eqnarray*} X_u(s)&=&x_0+\sum_t \frac{s^{|t|}}{|t|!} \alpha(t) t! A_u[\phi_v(t)] \delta_t\\ x(s)&=&x_0+\sum_t \frac{s^{|t|}}{|t|!} \alpha(t) t! \phi(t) \delta_t, \end{eqnarray*} where, as usually, \begin{eqnarray*} \phi_u(\parbox{3mm}{$\bullet$})&=&1\\ \phi_u(B^+(t_1\cdots t_k))&=&A_u[\phi_{u_1}(t_1)] \dots A_u[\phi_{u_k}(t_k)]\\ \phi(t)&=&B[\phi_u(t)]. \end{eqnarray*} Here $x(s)$ is the sum giving the bare quantity of interest. In the examples developed by Broadhurst and Kreimer \cite{Broadhurst}, the quantity of interest is \begin{eqnarray*} x(s)&=&\sum_t \frac{s^{|t|}}{|t|!} B_t, \end{eqnarray*} where $B_t$ is obtained recursively from given $B_n$ by \begin{eqnarray} B_\bullet &=& B_1 \nonumber\\ B_t &=& B_{|t|} \delta_{t_1} \delta_{t_2} \cdots \delta_{t_k} \quad\mathrm{when}\quad t=B^+(t_1,t_2,\cdots,t_k). \end{eqnarray} In the renormalization problems considered by Broadhurst and Kreimer, the $B_n$ are defined from a function $L(\delta)$ regular (and equal to 1) at the origin, by \begin{eqnarray*} B_n &=&\frac{L(n\epsilon)}{n\epsilon}. \end{eqnarray*} A pair of operators giving $\phi(t)=B_t$ can be defined as \begin{eqnarray*} A_u(X_v) &=& \frac{1}{\epsilon} \int_0^u L(\epsilon \frac{d}{dv}v)X_v, \quad B(X_v)=A_1(X_v). \end{eqnarray*} The quantity of interest $x(s)$ is then obtained by tensoring $A_u$ with the ``simple integral method'' to obtain $\phi(t)=B_t/t!$. The only thing that we need in the following is the action of $A_u$ on a monomial $v^{n-1}$ \begin{eqnarray} A_u(v^{n-1}) &=& \frac{1}{\epsilon} \int_0^u L(\epsilon \frac{d}{dv}v)v^{n-1}= \frac{1}{\epsilon} \int_0^u v^{n-1}dv L(n \epsilon )=B_n u^n. \label{actiondeAu} \end{eqnarray} \subsection{$S_R$, the ``renormalized antipode''} In ref.\cite{Kreimer}, Kreimer defines recursively a renormalized antipode\footnote{In Hopf algebra terms $S_R(\phi)(t)=-R[\phi(t)+m[(S_R\otimes Id)(\phi\otimes \phi)P_2\Delta(t)]$.} depending on a renormalization scheme $R$. We take as an example the toy model used by Kreimer, where $R[\phi]=\langle\phi\rangle$ is the projection of $\phi$ on the pole part of the Laurent series in $\epsilon$ inside the bracket. Following the results of section \ref{sectionantipode}, the Runge-Kutta method for $S_R(\phi)$ can be obtained from the Runge-Kutta method of $\phi$ by $A^S_u(X)=A_u(X)-\langle A_1(X)\rangle $, $B^S(X)=-\langle A_1(X)\rangle $. Working out the first examples using Eq.(\ref{actiondeAu}), we find, \begin{eqnarray*} \phi^S_u(\parbox{3mm}{$\bullet$})&=&1\\ \phi^S(\parbox{3mm}{$\bullet$})&=&\langle A_1(1) \rangle=\langle B_1 \rangle\\ \phi^S_u(\tdeux)&=&A_u(\phi_v(\parbox{3mm}{$\bullet$}))-\langle A_1(\phi_v(\parbox{3mm}{$\bullet$}))\rangle= A_u(1)-\langle A_1(1)\rangle=B_1 u- \langle B_1 \rangle\\ \phi^S(\tdeux)&=&-\langle A_1(\phi^S_u(\tdeux))\rangle= -\langle B_2 B_1 \rangle + \langle\langle B_1 \rangle B_1 \rangle. \end{eqnarray*} \subsection{Renormalized quantities} Finally, the renormalized quantities $x^R(s)$ are obtained from the convolution of $S_R(\phi)$ with $\phi$. To obtain the corresponding Runge-Kutta method, we use the results of section \ref{sectionconvolution}. However, the domain where $1\le u\le 2$ is not used, and the Runge-Kutta method for the renormalized quantity is $A^R_u(X)=A_u(X)-\langle A_1(X)\rangle $, $B^R(X)=A_1(X)-\langle A_1(X)\rangle $. It may seem surprising that such a simple equation encodes the full combinatorial complexity of renormalization. It is not even necessary to work examples out, because $A^R_u(X)=A^S_u(X)$ so that $\phi^R_u(t)=\phi^S_u(t)$, and the only difference comes from the action of $B^R$. For a real calculation of $x^R(s)$, we do not need $A^R_u$ and $B^R$ which give us $\phi(t)=\Gamma(t)$, but the tensor product of this method with the ``simple integral method'' to obtain $\phi(t)=t! \Gamma(t)$. In detail, the equation for the renormalized quantity $x^R(s)$ is \begin{eqnarray} X^R_{uu'}(s)&=&\frac{s}{\epsilon}\int_0^u dv \int_0^{u'} dv' L(\epsilon \partial_v v) e^{X_{vv'}(s)}- \langle \frac{s}{\epsilon}\int_0^1 dv \int_0^{u'} dv' L(\epsilon \partial_v v) e^{X_{vv'}(s)}\rangle \\ x^R(s)&=&\frac{s}{\epsilon}\int_0^1 dv \int_0^{1} dv' L(\epsilon \partial_v v) e^{X_{vv'}(s)}- \langle \frac{s}{\epsilon}\int_0^1 dv \int_0^{1} dv' L(\epsilon \partial_v v) e^{X_{vv'}(s)}\rangle. \label{Req1} \end{eqnarray} For a general renormalization scheme $R$, one replaces $\langle A_u(X) \rangle$ by $R[A_u(X)]$. Finally, Chen's lemma for renormalization schemes \cite{Kreimer} is obtained from Hairer and Wanner's theorem Eq.(\ref{Hairer2}). \section{Renormalization of Kreimer's toy model} In this section, we use Runge-Kutta methods to renormalized explicitly Kreimer's toy model for even functions $L(\epsilon)$. In \cite{Broadhurst}, remarkable properties of the renormalized sum of diagrams with ``Connes-Moscovici weights'' were noticed. \subsection{Equation for the renormalized quantity} The role of the sum over $u'$ in Eq.(\ref{Req1}) is to add a factor $1/t!$, as in section{\ref{sectionothersum}}. Therefore, the same reasoning can be used to show that $X^R_{uu'}(s)$ is in fact a function of $su'$ and we write $X^R_{uu'}(s)=X^R_{u}(su')$, which defines the function $X^R_{u}(s)$. The equation for $X^R_{u}(s)$ can be found from Eq.(\ref{Req1}) and the relation $X^R_{u}(s)=X^R_{us}(1)$ as \begin{eqnarray} X^R_{u}(s)&=&\frac{1}{\epsilon}\int_0^u dv \int_0^{s} ds' L(\epsilon \partial_v v) e^{X_{v}(s')}- \langle \frac{1}{\epsilon}\int_0^1 dv \int_0^{s} ds' L(\epsilon \partial_v v) e^{X_{v}(s')}\rangle \\ x^R(s)&=&\frac{1}{\epsilon}\int_0^1 dv \int_0^{s} ds' L(\epsilon \partial_v v) e^{X_{v}(s')}- \langle \frac{1}{\epsilon}\int_0^1 dv \int_0^{s} ds' L(\epsilon \partial_v v) e^{X_{v}(s')}\rangle. \label{Req2} \end{eqnarray} To solve this equation, we expand $X^R_{u}(s)$ in a power series over $u$: \begin{eqnarray*} X^R_{u}(s)&=&\sum_{n=0}^{\infty} a_n(s)u^n. \end{eqnarray*} A standard identity gives us \begin{eqnarray*} \exp(X^R_{u}(s))&=&\sum_{n=0}^{\infty} \lambda_n(a)u^n,\quad\mathrm{where}\\ \lambda_n(a)&=&\sum_{|\alpha|=n}\frac{a_1^{\alpha_1}\cdots a_n^{\alpha_n}} {{\alpha_1}!\cdots{\alpha_n}!}\quad\mathrm{with}\quad |\alpha|=a_1+2\alpha_2+\dots+n\alpha_n. \end{eqnarray*} $\lambda_n(a)$ depends on $s$ through its arguments $a_i(s)$. The sets of $\alpha_i$ for a given $n$ can be obtained from the partitions of $n$: $(\mu_1,\dots,\mu_n)$, where $\mu_1 \ge \cdots \ge \mu_n$ by $\alpha_n=\mu_n$, $\alpha_i=\mu_i-\mu_{i+1}$ for $i<n$. The first $\lambda_n(a)$ are \begin{eqnarray*} \lambda_0(a)&=&1\quad \lambda_1(a)=a_1\quad \lambda_2(a)=a_2+\frac{a_1^2}{2}\quad \lambda_3(a)=a_3+a_1a_2+\frac{a_1^3}{6}. \end{eqnarray*} \subsection{Solution of the equation} Introducing the series expansions for $X^R_{u}(s)$ and $\exp(X^R_{u}(s))$ into Eq.(\ref{Req2}) we obtain \begin{eqnarray*} \sum_{n=0}^{\infty} a_n(s)u^n&=&\sum_{n=0}^{\infty} B_{n+1} \int_0^s e^{a_0(s')} \lambda_n(a) ds' u^{n+1}- \langle\sum_{n=0}^{\infty} B_{n+1} \int_0^s e^{a_0(s')} \lambda_n(a) ds' \rangle \end{eqnarray*} or \begin{eqnarray} a_0(s)&=&-\langle\sum_{n=0}^{\infty} B_{n+1} \int_0^s e^{a_0(s')} \lambda_n(a) ds' \rangle\nonumber\\ a_n(s)&=&B_n \int_0^s e^{a_0(s')} \lambda_{n-1}(a) ds'\quad\mathrm{for}\quad n>0. \label{eqa0an} \end{eqnarray} To solve this equation, we need to go back to the equation for the bare quantity \begin{eqnarray} X^0_{u}(s)&=&\frac{1}{\epsilon}\int_0^u dv \int_0^{s} ds' L(\epsilon \partial_v v) e^{X^0_{v}(s')}. \label{eqbare} \end{eqnarray} Again $X^0_{u}(s)$ is a function of $su$, we define $X^0(s)=X^0_{s}(1)$ which satisfies \begin{eqnarray*} X^0(s)&=&\frac{1}{\epsilon}\int_0^s \frac{du}{u} \int_0^{u} dv L(\epsilon \partial_v v) e^{X^0(v)}. \end{eqnarray*} The solution of this equation is given by the B-series \begin{eqnarray} X^0(s)&=&\sum_n {\bar{\alpha}}_n s^n\quad\mathrm{with}\quad {\bar{\alpha}}_n = \sum_{|t|=n} \frac{\alpha(t)B_t}{|t|!}. \label{bareeq} \end{eqnarray} On the other hand, we can also expand $e^{X^0(v)}$ using the functions $\lambda_n(\bar{a})$. Identifying both sides of Eq.(\ref{bareeq}), we obtain the relation \begin{eqnarray} {\bar{a}}_n&=&\frac{B_n}{n}\lambda_{n-1}(\bar{a}).\label{abar} \end{eqnarray} With this identity, we can now prove that, for the renormalized quantities, \begin{eqnarray} a_n(s)&=&(g(s))^n{\bar{a}}_n,\quad\mathrm{where}\quad g(s)=\int_0^s\exp(a_0(s'))ds'. \label{aid} \end{eqnarray} Since $\lambda_0(a)=1$ and ${\bar{a}}_n=B_1$, this equation is true for $n=1$, from Eq.(\ref{eqa0an}). If Eq.(\ref{aid}) is true up to $n-1$, then $\lambda_{n-1}(a)=(g(s))^{n-1}\lambda_{n-1}(\bar{a})$ and the derivative of Eq.(\ref{eqa0an}) gives us \begin{eqnarray*} a'_n(s)&=&B_n e^{a_0(s)} \lambda_{n-1}(a)= B_n g'(s) (g(s))^{n-1} \lambda_{n-1}(\bar{a})=n(g(s))^{n-1} \bar{a}_n, \end{eqnarray*} by Eq.(\ref{abar}). Integrating this equation with the condition $a_n(s)=0$ gives Eq.(\ref{aid}) at level $n$. By this we have proved that the flow for the renormalized quantity is a reparametrization of the flow for the bare quantity: $X^R_u(s)=a_0(s)+X^0(u g(s))$ and $X^R(s)=a_0(s)+X^0(g(s))$. To determine $a_0(s)$ we proceed step by step. In Eq.(\ref{bareeq}) we expand $L(\epsilon \partial_v v)$ over $\epsilon$. The first term is just 1, and we obtain Eq.(\ref{doublefact}) with the solution $x(s)=-2\log(1-s/(2\epsilon))$. For the renormalized quantity, the most singular term becomes $X^0(g(s))=-2\log(1-g(s)/(2\epsilon))$. Since $X^R(s)$ is regular, this singular term must be compensated by a corresponding term in $a_0(s)$. By equating the most singular terms we obtain $a_0(s)=-2\log(1-g(s)/(2\epsilon))$. We know from Eq.(\ref{aid}) that $a_0(s)=\log(g'(s))$, and we obtain the most singular terms as the solution of $g'(s)=1/(1-g(s)/(2\epsilon))^2$, which is: \begin{eqnarray*} g(s)&=&\frac{s}{1+\frac{s}{2\epsilon}}\\ a_0(s)&=& -2\log(1+\frac{s}{2\epsilon}). \end{eqnarray*} By expanding $a_0(s)$ as a series in $s$, we obtain the most singular term observed in \cite{Broadhurst} and proved in \cite{Kreimer}. One notices that the singularity of the non-pertubative term $a_0(s)$ is logarithmic, and much smoother than the singularities coming from the expansion over $s$ (i.e. the perturbative expression). \subsection{Differential equation for the finite part} In general, one should proceed now with the next singular term. To obtain it we denote $Y(s)=X^0(g(s))$, this change of variable gives the equation for $Y(s)$: \begin{eqnarray*} Y(s)&=&\frac{1}{\epsilon}\int_0^s \frac{g'(u)du}{g(u)} \int_0^{u} dv g'(v) L(\epsilon+\epsilon \frac{g(v)}{g'(v)} \partial_v ) e^{Y(v)}. \end{eqnarray*} Now we can write $Y(s)=X^R(s)-a_0(s)$, and notice that the term $-a_0(s)$ on the left-hand side is compensated by a term on the right-hand side where $L=1$ and $\exp(X^R(s))=1$. We obtain the equation for $X^R(s)$: \begin{eqnarray*} X^R(s)&=&\frac{1}{\epsilon}\int_0^s \frac{du}{u(1+\frac{u}{2\epsilon})} \int_0^{u} dv\left[ \frac{1}{(1+\frac{v}{2\epsilon})^2} L(\epsilon \partial_v v+\frac{v^2}{2} \partial_v){(1+\frac{v}{2\epsilon})}^2 e^{X^R(v)} -1\right]. \end{eqnarray*} The nice aspect of the previous equation is that it seems to have a limit as $\epsilon$ goes to zero. In fact, it has a limit when $L$ is even, as we shall show now. Writing $\bar{X}(s)=\lim_{\epsilon\rightarrow 0}X^R(s)$, and taking the limit $\epsilon\rightarrow 0$ in the previous equation, we obtain \begin{eqnarray*} \bar{X}(s)&=&2\int_0^s \frac{du}{u^2} \int_0^{u} dv\left[ \frac{1}{v^2} L(\frac{v^2}{2} \partial_v )v^2 e^{\bar{X}(v)} -1\right], \end{eqnarray*} or, in differential form: \begin{eqnarray*} \frac{1}{2}(s^2\bar{X}'(s))'&=&\frac{1}{s^2} L(\frac{s^2}{2}\frac{d}{ds})s^2 e^{\bar{X}(s)} -1. \end{eqnarray*} If $\bar{X}(s)$ and $ L(\delta)$ are expanded as \begin{eqnarray} \bar{X}(s)&=& \sum_{n=1}^\infty b_n s^n\quad\mathrm{and}\quad L(\delta)= 1+\sum_{n=1}^\infty L_n \delta^n,\quad\mathrm{so that} \nonumber\\ L(\frac{s^2}{2}\frac{d}{ds})&=& 1+\sum_{n=1}^\infty L_n (\frac{s^2}{2}\frac{d}{ds})^n, \label{renoreq} \end{eqnarray} we obtain the following relation for the term in $s$: $b_1s=(b_1+L_1/2)s$. If $L_1$ is not zero, we obtain a contradition and must proceed with the withdrawal of divergences. For simplicity, we shall assume that $L_1=0$. Then $b_1$ becomes a free parameter of $\bar{X}(s)$. All terms $b_n$ with $n>1$ can now be determined from $b_1$ and $L_n$ ($n>1$). All terms are regular. In \cite{Broadhurst}, the function $L(\delta)$ was taken even. Then $L_1=0$, and their results correspond to $b_1=0$. Broadhurst and Kreimer have also used a function $L(\epsilon,\delta)$. The present treatment can be applied to this more general situation, with the only change that \begin{eqnarray*} L_n=n!\lim_{\epsilon\rightarrow 0}\lim_{\delta\rightarrow 0} \frac{d^n}{d\delta^n} L(\epsilon,\delta). \end{eqnarray*} Clearly, Eq.(\ref{renoreq}) is much faster to solve than computing the sum over trees. For instance, the expansion could be calculated up to 20 loops (i.e. $b_{20}$) within a few seconds with a computer. \subsection{Alternative point of view} There is an alternative way to solve Eq.(\ref{eqbare}) for the bare quantity. We define a function $f(s)$ from $L(\delta)$ by \begin{eqnarray*} f(s)&=& \sum_{n=0}^\infty \frac{L(n\epsilon+\epsilon)}{n!} s^n= L(\epsilon \frac{d}{ds}s) e^s. \end{eqnarray*} A relation between $f(s)$ and $L(\delta)$ can also be established through the Mellin transforms of $f$ and $L$ as $M(f)(z)=M(L)(\epsilon-\epsilon z)\Gamma(z)$. With $f(s)$ we can write the equation for the bare quantity as \begin{eqnarray} X^0(s)&=&\frac{1}{\epsilon}\int_0^s \frac{du}{u} \int_0^{u} dv f(X^0(v)). \label{eqavecf} \end{eqnarray} Alternatively, one can go from $f$ to $L$ and consider the results of the toy model as a method to renormalize equations of the type (\ref{eqavecf}). \section{n-dimensional problems} For applications to classical field theory, we need to develop Runge-Kutta methods for the n-dimensional analogue of the flow equation: non-linear partial differential equations. The purpose of the present section is to indicate how B-series can be used for this case\footnote{Kreimer was independently aware of the possibility to use B-series for non-linear partial differential equations.}. The method apply to equations of the form $L\psi(\mathbf{r})= F[\psi(\mathbf{r})]$, where $L$ is a differential operator (e.g. the nonlinear Schr\"odinger equation $\Delta\psi=\psi^3$). \subsection{Formulation} We need two starting elements: a function $\psi_0(\mathbf{r})$ which is the solution of $L\psi_0(\mathbf{r})=0$, and a Green function $G(\mathbf{r},\mathbf{r}')$, that is a solution of the equation $L_rG(\mathbf{r},\mathbf{r}')=\delta(\mathbf{r}-\mathbf{r}')$, with given boundary conditions. The function $\psi_0(\mathbf{r})$ will play the role of an initial value, and the Green function will decide in which ``direction'' you move from the initial value. It will also state, in some sense, the boundary conditions of the solution $\psi(\mathbf{r})$. Using these two functions, the differential equation $L\psi(\mathbf{r})=F[\psi(\mathbf{r})]$ is transformed into $\psi(\mathbf{r})=\psi_0(\mathbf{r})+\int d\mathbf{r}' G(\mathbf{r},\mathbf{r}')F[\psi(\mathbf{r}')]$. The action of $L$ enables us to go from the second to the first equation. The combinatorics is the same as for the standard Runge-Kutta method, and the result is \begin{eqnarray} \psi(\mathbf{r})&=&\psi_0(\mathbf{r})+\sum_t \frac{\alpha(t) t!}{|t|!} \int d\mathbf{r}' G(\mathbf{r},\mathbf{r}') \phi_{r'}(t),\label{Bseries_n} \end{eqnarray} where $\phi_{r}(t)$ is defined recursively by \begin{eqnarray} \phi_{r}(\parbox{3mm}{$\bullet$})&=&F[\psi_0(\mathbf{r})]\nonumber\\ \phi_{r}(B^+(t_1\cdots t_k))&=&F^{(k)}[\psi_0(\mathbf{r})] \int dr_1 G(\mathbf{r},\mathbf{r}_1)\phi_{r_1}(t_1) \dots \int dr_k G(\mathbf{r},\mathbf{r}_k)\phi_{r_k}(t_k). \label{Bphi_n} \end{eqnarray} If $\psi$ is a vector field, the solution is the same, and equations (\ref{Bphi_n}) get indices: \begin{eqnarray*} \phi^i_{r}(\parbox{3mm}{$\bullet$})&=&f^i[\psi_0(\mathbf{r})]\\ \phi^i_{r}(B^+(t_1\cdots t_k))&=&F^i_{j_1\dots j_k}[\psi_0(\mathbf{r})] \int dr_1 G^{j_1}_{j'_1}(\mathbf{r},\mathbf{r}_1)\phi^{j'_1}_{r_1}(t_1) \dots \int dr_k G^{j_k}_{j'_k}(\mathbf{r},\mathbf{r}_k)\phi^{j'_k}_{r_k}(t_k), \end{eqnarray*} where $G^i_j(\mathbf{r},\mathbf{r}')$ is a component of the matrix Green function. In the previous sections, the series (\ref{Bseries}) was written as a function of $\phi(t)$ (describing the effect of the Runge-Kutta method ($a$,$b$)) and $\delta_t$ (describing the effect of the function $F[x]$). In the present case, this separation is no longer possible, and $\phi(t)$ combines both pieces of information. \subsection{Examples} In this section, equation (\ref{Bseries_n}) is applied to the one-dimensional problem and to the Schr\"odinger equation. \subsubsection{The one-dimensional case} It is instructive to observe how the one-dimensional case is obtained from Eq.(\ref{Bseries_n}). The differential operator is $L=d/ds$, so the initial function $\psi_0(s)$ must satisfy $d/ds\psi_0(s)=0$: $\psi_0(s)$ is a constant that we write $x_0$. For the Green function $G(s,s')$, we have the equation $LG(s,s')=\delta(s-s')$, so $G(s,s')=\theta(s-s') +C(s')$, where $\theta(s)$ is the step function and $C(s')$ a function of $s'$. To determine $C(s')$, we note that, in the ``simple integral method'', there is an integral from $s_0$ to $s$. From the Green function $G(s,s')=\theta(s-s') -\theta(s_0-s')$, we obtain \begin{eqnarray*} \int_{-\infty}^{\infty} G(s,s') f(s') ds'&=& \int_{s_0}^{s} f(s') ds' \end{eqnarray*} which is the required expression. Now, the role of $\psi_0$ and the Green function is clear for the one-dimensional case: $\psi_0$ gives the initial value $x_0$ and $G$ specifies (among other things) the starting point $s_0$. To complete the derivation of the one-dimensional case, we note that $\psi_0(s)=x_0$ does not depend on $s$, so the terms $F^{(k)}[\psi_0(s)]=F^{(k)}[x_0]$ are independent of $s$ and can be grouped together to build $\delta_t$ as in (\ref{defdeltat}). On the other hand, the integration over Green functions build up $(s-s_0)^{|t|}/t!$ and we obtain Eq.(\ref{Butcherflow}). \subsubsection{The Schr\"odinger equation I} If we write the Schr\"odinger equation as $(E+\Delta)\psi(\mathbf{r})=V(\mathbf{r})\psi(\mathbf{r})$, we can apply Eq.(\ref{Bseries_n}) with $F[\psi]=V(\mathbf{r})\psi$. We take for $\phi_0(\mathbf{r})$ a solution of $(E+\Delta)\psi_0(\mathbf{r})=0$ and for $G(\mathbf{r},\mathbf{r}')$ the scattering Green function (e.g. $G(\mathbf{r}-\mathbf{r}')=-e^{i\sqrt{E}|\mathbf{r}-\mathbf{r}'|}/(4\pi |\mathbf{r},\mathbf{r}'|)$ in three dimensions). The calculation of $\phi(t)$ is straightforward because, in a such a linear problem, $F^{(k)}=0$ for $k>1$. Hence, the only rooted trees that survive are those with one branch. For these trees $\alpha(t)=1$ and $t!=|t|!$ and we obtain \begin{eqnarray*} \psi(\mathbf{r})=\psi_0(\mathbf{r})+\int d\mathbf{r}_1 G(\mathbf{r},\mathbf{r}_1) V(\mathbf{r}_1)\psi_0(\mathbf{r}_1)+ \int d\mathbf{r}_1 d\mathbf{r}_2 G(\mathbf{r},\mathbf{r}_1) V(\mathbf{r}_1) G(\mathbf{r}_1,\mathbf{r}_2) V(\mathbf{r}_2)\psi_0(\mathbf{r}_2)+\cdots \end{eqnarray*} where we recognize the Born expansion of the Lippmann-Schwinger equation. \subsubsection{The Schr\"odinger equation II} We can also treat the Schr\"odinger equation in an alternative way as the system of equations: \begin{eqnarray*} (E+\Delta)\psi(\mathbf{r})&=& V(\rho)\psi(\mathbf{r}) \\ \frac{\partial \rho_i}{\partial r_j} &=& \delta_{ij}. \end{eqnarray*} This is a matrix differential equation. We give index 0 to the first line, and index $i$ (running from 1 to the dimension of space) to the other lines, called the space lines. The purpose of the space lines is just to ensure that $\rho=\mathbf{r}$. This is a standard trick to take the $\mathbf{r}$ dependence of $V$ into account in the expansion (see e.g. \cite{Hairer} p.143). As initial value we take $\psi_0(\mathbf{r})$ and $\rho_0=0$, the matrix Green function is diagonal and it is equal to the scattering wave function for line 0 and to $\theta(r_i-r'_i)-\theta(-r'_i)$ for line $i$. For $\phi_r(\parbox{3mm}{$\bullet$})$, the zero-th component is $V(0)\psi_0(\mathbf{r})$ and the space components are 1, for all the other trees, the space components are 0 and the zero-th component of the simplest tree is \begin{eqnarray*} \phi_r(\tdeux)&=&V(0)^2 \int d\mathbf{r}' G(\mathbf{r},\mathbf{r}') \psi_0(\mathbf{r}')+ \sum_i r_i\partial_i V(0) \psi_0(\mathbf{r})\\ \phi_r(\ttroisun)&=&V(0)^3 \int d\mathbf{r}_1 G(\mathbf{r},\mathbf{r}_1) + V(0) \int d\mathbf{r}' G(\mathbf{r},\mathbf{r}') \sum_i r'_i\partial_i V(0) \psi_0(\mathbf{r}')\\ \phi_r(\ttroisdeux)&=&2V(0)\sum_i r_i\partial_i V(0) \int d\mathbf{r}' G(\mathbf{r},\mathbf{r}') \psi_0(\mathbf{r}')+ \sum_{ij} r_i r_j\partial_i\partial_j V(0). \end{eqnarray*} The expressions become more and more complex, but their derivation is made systematic by the recurrence relation. \section{Conclusion} Butcher's approach to Runge-Kutta methods was applied to some simple renormalization problems. Since Cayley, it is clear that the ART is ideally suited to treat differentials. This was confirmed here by presenting a B-series solution of a class of non-linear partial differential equations. The recursive nature of B-series make them computationally efficient: $\phi_u(t)$ can be obtained by a simple operation from the $\phi_u(t')$ of smaller order $t'$. This is why B-series can be automated and implemented in a computer. Butcher's approach has still much to offer. In the numerical analysis literature, B-series have been generalized to treat flow equations on Lie groups. The main change \cite{Munthe98} is to replace the algebra of rooted trees by the algebra of planar trees (also called ordered trees \cite{Owren}). The elementary differentials get then a ``quantized calculus'' flavor, especially in the definition given Munthe-Kaas \cite{Munthe95} in terms of commutators with the vector field $F=f^i\partial_i$ (see also Ginocchio). Using this generalized ART, extended work has been carried out recently for the numerical solution of differential equations on Lie groups (see Ref.\cite{Munthe98,Owren} and the web site {\tt http://www.math.ntnu.no/num/synode}). B-series have been generalized in other directions, e.g. stochastic differential equations \cite{Komori} and differential equations of the type $dy/ds=f(y,z)$, $g(y,z)=0$, which are called differential algebraic equations \cite{HairerII}. It is our hope that Butcher's approach can be applied to quantum field theory. \section{Acknowledgements} It is my great pleasure to thank Dirk Kreimer and Alain Connes for interest, encouragement and discussions. \section{Appendix} For further reference, the action of the coproduct and the antipode on the first few trees are given here. \subsection{Coproduct} \begin{eqnarray*} \Delta 1 &=& 1 \otimes 1 \\ \Delta \parbox{3mm}{$\bullet$} &=& \parbox{3mm}{$\bullet$}\otimes 1 + 1 \otimes \parbox{3mm}{$\bullet$} \\ \Delta \tdeux &=& \tdeux\otimes 1 + 1 \otimes \tdeux+ \parbox{3mm}{$\bullet$}\otimes \parbox{3mm}{$\bullet$} \\ \Delta \ttroisun &=& \ttroisun\otimes 1 + 1 \otimes \ttroisun + \tdeux\otimes \parbox{3mm}{$\bullet$}+ \parbox{3mm}{$\bullet$}\otimes \tdeux \\ \Delta \ttroisdeux &=& \ttroisdeux\otimes 1 + 1 \otimes \ttroisdeux + \parbox{3mm}{$\bullet$}\tun\otimes \parbox{3mm}{$\bullet$}+2 \parbox{3mm}{$\bullet$}\otimes \tdeux \\ \Delta \tquatreun &=& \tquatreun\otimes 1 + 1 \otimes \tquatreun + \parbox{3mm}{$\bullet$}\otimes \ttroisun+ \ttroisun\otimes \parbox{3mm}{$\bullet$}+ \tdeux\otimes \tdeux \\ \Delta \tquatredeux &=& \tquatredeux\otimes 1 + 1 \otimes \tquatredeux + 2 \parbox{3mm}{$\bullet$}\otimes \ttroisun+ \ttroisdeux\otimes \parbox{3mm}{$\bullet$}+ \parbox{3mm}{$\bullet$}\tun\otimes \tdeux \\ \Delta \tquatretrois &=& \tquatretrois\otimes 1 + 1 \otimes \tquatretrois + \parbox{3mm}{$\bullet$}\otimes \ttroisun+ \parbox{3mm}{$\bullet$}\otimes \ttroisdeux+ \tdeux\otimes \tdeux+ \parbox{3mm}{$\bullet$}\tun\otimes \tdeux+ \parbox{3mm}{$\bullet$}\tdeux\otimes\parbox{3mm}{$\bullet$} \\ \Delta \tquatrequatre &=& \tquatrequatre\otimes 1 + 1 \otimes \tquatrequatre + 3\parbox{3mm}{$\bullet$}\otimes \ttroisdeux+ 3\parbox{3mm}{$\bullet$}\tun\otimes \tdeux+ \parbox{3mm}{$\bullet$}\tun\parbox{3mm}{$\bullet$}\otimes \parbox{3mm}{$\bullet$} \end{eqnarray*} \subsection{Antipode} \begin{eqnarray*} S(1) &=& 1 \\ S(\parbox{3mm}{$\bullet$}) &=& -\parbox{3mm}{$\bullet$} \\ S\left(\tdeux\right) &=& -\tdeux+ \parbox{3mm}{$\bullet$} \parbox{3mm}{$\bullet$} \\ S\left( \ttroisun\right) &=& -\ttroisun +2 \parbox{3mm}{$\bullet$}\tdeux -\parbox{3mm}{$\bullet$}\tun\parbox{3mm}{$\bullet$} \\ S\left( \ttroisdeux\right) &=& -\ttroisdeux +2 \parbox{3mm}{$\bullet$}\tdeux -\parbox{3mm}{$\bullet$}\tun\parbox{3mm}{$\bullet$} \\ S\left( \tquatreun\right) &=& -\tquatreun+2 \parbox{3mm}{$\bullet$} \ttroisun+ \tdeux \tdeux-3\parbox{3mm}{$\bullet$}\tun\tdeux+ \parbox{3mm}{$\bullet$}\tun\parbox{3mm}{$\bullet$}\tun \\ S\left( \tquatredeux\right) &=& -\tquatredeux+2 \parbox{3mm}{$\bullet$} \ttroisun+ \parbox{3mm}{$\bullet$} \ttroisdeux-3\parbox{3mm}{$\bullet$}\tun\tdeux+ \parbox{3mm}{$\bullet$}\tun\parbox{3mm}{$\bullet$}\tun \\ S\left( \tquatretrois\right) &=& -\,\tquatretrois+\parbox{3mm}{$\bullet$} \ttroisun+ \parbox{3mm}{$\bullet$} \ttroisdeux+ \tdeux \tdeux-3\parbox{3mm}{$\bullet$}\tun\tdeux+ \parbox{3mm}{$\bullet$}\tun\parbox{3mm}{$\bullet$}\tun \\ S\left( \tquatrequatre\right) &=& -\tquatrequatre+ 3\parbox{3mm}{$\bullet$} \ttroisdeux -3\parbox{3mm}{$\bullet$}\tun\tdeux+ \parbox{3mm}{$\bullet$}\tun\parbox{3mm}{$\bullet$}\tun \end{eqnarray*} \end{fmffile}
\section{Introduction} It is now well known that the statistical analysis of weak lensing effects on background galaxies due to foreground large scale structure can be used as a probe of cosmological parameters, such as the matter density and the cosmological constant, and the projected mass distribution of the Universe (e.g., Bernadeau et al. 1997; Blandford et al. 1991; Jain \& Seljak 1997; Kaiser 1998; Miralda-Escud\'e 1991; Schneider et al. 1998). Given that weak gravitational lensing results from the projected mass distribution, the statistical properties of weak lensing, such as the two point function of the shear distribution, reflects certain aspects associated with the projected matter distribution. With growing interest in weak gravitational lensing surveys, several studies have now explored the accuracy to which conventional cosmological parameters, such as the mass density of the Universe and the cosmological constant can be determined (e.g., Bartelmann \& Schneider 1999; van Waerbeke et al. 1999). Beyond primary cosmological parameters, such as the mass density, the projected matter distribution is also affected by the presence of neutrinos with non-zero masses. For example, when non-zero mass neutrinos are present, a strong suppression of power in the mass distribution occurs at scales below the time-dependent free streaming scale (see, e.g., Hu \& Eisenstein 1998). The detection of such suppression, say in the density power spectrum, allows a direct measurement of the neutrino mass, in contrast to various particle physics based neutrino experiments which only allow measurements of mass differences, or splittings, between different neutrino species (e.g., Super Kamiokande experiment; Fukuda et al. 1998). The direct astrophysical probes of neutrino masses include time-of-flight from a core-collapse supernova (e.g., Totani 1998; Beacom \& Vogel 1998; see, Beacom 1999 for a review) and large scale structure power spectrum. The suppression of power at small scales due to neutrinos can easily be investigated with the galaxy power spectrum from wide-field redshift surveys, such as the Sloan Digital Sky Survey (SDSS\footnote{http://www.sdss.org/}; e.g., Hu et al. 1997), however, such measurements are subjected to unknown biases between galaxy and matter distribution and its evolution with redshift. Therefore, an understanding of bias and its evolution may first be necessary before making a reliable measurement of neutrino mass using galaxy power spectra. On the contrary, a measurement of the power spectrum unaffected by such effects allows a strong possibility to measure the neutrino mass. Such a possibility should soon be available with weak gravitational lensing surveys through the measured weak lensing power spectrum directly, which probes the matter power spectrum through a convolution of the redshift distribution of sources and distances. Thus, it is expected that the weak lensing power spectrum can also allow a determination of the suppression due to neutrinos, and thus, a direct measurement of the neutrino mass. In addition to neutrino mass, weak lensing also allows determination of several cosmological parameters, including matter density ($\Omega_m$) and the cosmological constant ($\Omega_\Lambda$). However, there are large number of cosmological probes that essentially measure these parameters. For example, luminosity distance measurements to Type Ia supernovae at high redshifts and gravitational lensing statistics allow the determination of $\Omega_m$ and $\Omega_\Lambda$ (e.g., Cooray et al. 1998; Perlmutter et al. 1998; Riess et al. 1998), or rather $\Omega_m-\Omega_\Lambda$, while the location of the first Doppler peak in the cosmic microwave background (CMB) power spectrum allows a determination of these two quantities in the orthogonal direction ($\Omega_m+\Omega_\Lambda$; White 1998). When combined (e.g., Lineweaver 1998), these two parameters can be known to a high accuracy reaching a level of $\sim$ 5\%, when the expected measurements on SNe and CMB experiments over the next decade, such as MAP, are considered (e.g., Tegmark et al. 1998). Since these measurements are not sensitive to neutrinos, therefore, it is necessary that one returns to probes which are strongly sensitive to neutrinos to obtain important cosmological information on their presence. This is the primary motivation of this paper: weak lensing is highly suitable for a neutrino mass measurement when compared to various other probes of cosmology, including CMB. The direct measurement of galaxy power spectrum also allows a measurement of neutrino mass, as has been discussed in Hu et al. (1997), however as discussed above, such a measurement can be contaminated by bias and its evolution. The other motivation for this paper comes through the cosmological importance of neutrinos (see, Ma 1999 for a recent review on this subject). Current upper bounds on neutrino masses range from 23 eV based on SN 1987A neutrino arrival time delays to 4.4 eV using recent oscillation experiments, assuming that 3 degenerate neutrino species are present (Vernon Barger, priv. communication; Fogli et al. 1997. Recently, Croft et al. (1999) determined that $m_\nu < 5.5$ eV at the 95\% level using the Ly$\alpha$ forest for all $\Omega_m$ values and $m_\nu < 2.4 (\Omega_m/0.17 - 1)$ eV for $0.2 \la \Omega_m \la 0.5$ (95\% confidence). A rather definite upper limit on neutrino mass is 94 eV, which is the mass limit to produce a normalized cosmological mass density of 1, while according to Ma (1999), a rather conservative cosmological limit on neutrino mass presently is $\sim$ 5 eV. However, apart from these limits, several studies still suggest the possibility that the neutrino mass can be as high as 15 eV (e.g., Shi \& Fuller 1999), therefore, it is safe to say that neutrino mass or limit on its mass is not strongly constrained. The neutrino mass is one of the important cosmological parameters, and thus, it is necessary that suitable probes which allow this measurement, beyond mass splitting measurements allowed by particle physics experiments, be studied. In this paper, we explore the possibility for a neutrino mass measurement with weak lensing surveys and suggest that weak lensing can be used as a strong probe of the neutrino mass, provided that one has adequate knowledge on the uncertainties of basic cosmological parameters will other techniques. In a recent paper, Hu \& Tegmark (1999) explored the full parameter space of wide-field weak lensing surveys combined with future cosmic microwave background (CMB) satellites. A recent review on weak lensing could be found in Mellier (1998). In Sect.~2, we discuss the effect of neutrinos in the weak lensing convergence power spectrum and calculate accuracies to which the neutrino mass can be determined. We follow the conventions that the Hubble constant, $H_0$, is 100\,$h$\ km~s$^{-1}$~Mpc$^{-1}$ and $\Omega_i$ is the fraction of the critical density contributed by the $i$th energy component: $b$ baryons, $\nu$ neutrinos, $m$ all matter species (including baryons and neutrinos) and $\Lambda$ cosmological constant. \section{Weak Lensing Power Spectrum} \subsection{Effective Convergence Power Spectrum} Following Kaiser (1998) and Jain \& Seljak (1997), we can write the power spectrum of convergence due to weak gravitational lensing as: \begin{equation} P_\kappa(l) = l^4 \int d\chi \frac{g^2(\chi)}{r^6(\chi)} P_\Phi\left(\frac{l}{r(\chi)},\chi\right), \end{equation} where $\chi$ is the radial comoving distance related to redshift $z$ through: \begin{equation} \chi(z) = \frac{c}{H_0}\int_{0}^{z} dz' \left[ \Omega_m (1+z')^3 + \Omega_k (1+z')^2 + \Omega_\Lambda \right]^{-1/2}, \end{equation} and $r(\chi)$ is the comoving angular diameter distance written as $r(\chi) = 1/\sqrt{-K} \sin \sqrt{-K}\chi,\chi, 1/\sqrt{K}\sinh \sqrt{K}\chi$ for closed, flat and open models respectively with $K = (1-\Omega_{\rm tot})H_0^2/c^2$. In Eq.~1, $P_\Phi(k=\frac{l}{r(\chi)},\chi)$ is the time-dependent three dimensional power spectrum of the Newtonian potential which is related to the density power spectrum, $P_\delta(k)$, through the Poisson's equation (e.g., Eq.~2.6 of Schneider et al. 1998) and $g(\chi)$ weights the background source distribution by the lensing probability: \begin{equation} g(\chi) = r(\chi) \int_{\chi}^{\chi_H} \frac{r(\chi'-\chi)}{r(\chi')} W_\chi(\chi') d\chi'. \end{equation} Here, $\chi_H$ is the comoving distance to the horizon. Following Kaiser (1998) and Hu \& Tegmark (1999), we can write the expected uncertainties in the weak lensing convergence power spectrum as: \begin{equation} \sigma(P_\kappa)(l) = \sqrt{\frac{2}{(2\l+1)f_{\rm sky}}}\left( P_\kappa(l) + \frac{\langle \gamma^2 \rangle}{n_{\rm mag}}\right), \end{equation} where $f_{\rm sky}$ is the fraction of the sky covered by a survey, $\sqrt{\langle \gamma^2 \rangle} \sim 0.4$ is the intrinsic non-zero ellipticity of background galaxies and $n_{\rm mag}$ is the surface density of galaxies down to the magnitude limit of the survey. Thus, Eq.~5, accounts for three sources of noise in the weak lensing power spectrum: cosmic variance, shot-noise in the ellipticity measurements and the number of galaxies available to make such measurements from which the weak lensing properties are derived (see, e.g., Schneider et al. 1998 and Kaiser 1998 for further details). We take $n_{\rm mag}$ to be $6.5 \times 10^{8}$ sr$^{-1}$ down to R magnitude of 25 and $4 \times 10^{9}$ sr$^{-1}$ down to R of 27, which were determined based on galaxy number counts of deep surveys such as the Hubble Deep Field. Following Schneider et al. (1998), we parameterize the source distribution, $W_\chi(\chi)$, as a function of redshift, $W_z(z)$: \begin{equation} W_z(z) = \frac{\beta}{\Gamma\left[\frac{1+\alpha}{\beta}\right] z_0} \left(\frac{z}{z_0}\right)^\alpha \exp\left[-\left(\frac{z}{z_0}\right)^\beta\right]. \end{equation} Such a distribution has been observed to provide a good fit to the observed redshift distribution of galaxies (e.g., Smail et al. 1995). \subsection{Linear and Nonlinear Power Spectra} Since we are considering non-zero mass neutrinos, it is necessary that both a linear and a nonlinear power spectrum which takes in to account for such neutrinos be considered. In order to obtain the linear power spectrum, we follow Hu \& Eisenstein (1998) and Eisenstein \& Hu (1999) and consider the MDM (mixed dark matter) transfer and growth functions appropriate for massive neutrinos as well as baryons. We use fitting formulae presented therein which agree with numerical calculations at a level of 1\%. Both neutrinos and baryons affect the standard power spectrum by suppressing power at small scales below the free-streaming length. The small scale suppression due to neutrinos can be written as: \begin{equation} \left(\frac{\Delta P}{P}\right) \sim -8\frac{\Omega_\nu}{\Omega_m} \sim -0.8\left(\frac{m_\nu}{{\rm 1\,eV}}\right)\left(\frac{0.1N}{\Omega_m h^2}\right), \end{equation} where $N$ is the number of degenerate neutrinos. Assuming the standard model for neutrinos with a temperature $(4/11)^{1/3}$ that of the CMB, we can write $\Omega_\nu$ based on neutrino mass, $m_\nu$ (in eV), and the number of degenerate neutrino species, $N$, as $\Omega_\nu = N(m_\nu/94)h^{-2}$. We assume integer number of neutrino species that can amount up to three. The suppression of power is proportional to the ratio of hot matter density of neutrinos to cold matter density; in low $\Omega_m$ cosmological models, currently preferred by observations, the suppression of power is much larger than in an Einstein-de Sitter universe with the same amount of neutrinos. In fact, for low $\Omega_m$ models, massive neutrinos of mass $\sim$ 1 eV, contribute to a 100\% suppression of power compared with no neutrinos. In addition to the linear power spectrum, given the time-dependence, it is necessary that the non-linear evolution of the density power spectrum be fully taken into account when calculating the convergence power spectrum given in Eq.~1. The importance of the non linear evolution of the power spectrum on weak lensing statistics was first discussed in Jain \& Seljak (1997) for standard $\Lambda$CDM cosmological models involving $\Omega_m$ and $\Omega_\Lambda$. There are several approaches to obtain the nonlinear evolution, however, for analytical calculations, fitting functions are strongly preferred over detailed numerical work. In Peacock \& Dodds (1996), the evolved density power spectrum was related to the linear power spectrum through a function $F(x)$, where $x$ was calibrated against numerical simulations in standard CDM models. According to Peacock \& Dodds (1996), the nonlinear power spectrum $P_\delta$ is related to the linear power spectrum, $P_\delta^L(k_L)$, through: $k^3P_\delta(k)/(2\pi^2) = F\left[k_L^3P_\delta^L(k_L)/(2\pi^2)\right]$ where $k_L = \left[1+ k^3P_\delta(k)/(2\pi^2)\right]^{-1/3}k$. We refer the reader to Peacock \& Dodds (1996) for the functional form of $F(x)$. Since we are now allowing for the presence of massive neutrinos, as well as baryons, it is necessary that we consider whether the fitting function given in Peacock \& Dodds (1996) is reliable for the present calculation as these two species were not included in their simulations; Smith et al. (1998) compared the Peacock \& Dodds (1996) formulation against MDM numerical simulations and suggested a possible agreement between the two when spectral index for Peacock \& Dodds (1996) fitting formula was calculated using MDM power spectrum. However, recently, Ma (1998) suggested that this agreement was only due to poor resolution of numerical simulations used in Smith et al. (1998). According to Ma (1998), Peacock \& Dodds (1996) formulation disagrees with numerical data at the level of 10\% to 50\%. Therefore, instead of the Peacock \& Dodds (1996) approach, we use the fitting function given in Ma (1998) in the present calculation which was now shown to agree with numerical simulations at a level of 3\% to 10\% for $k \la 10\; h$ Mpc$^{-1}$ out to a redshift of $\sim 4$. For higher scales and redshifts, agreement is only reached at a level of 15\% against numerical simulations. For the present calculation involving a redshift distribution that peaks at redshifts lower than 4 with scales of interest lower than 10 $h^{-1}$ Mpc$^{-1}$, the fitted formulation is reasonably adequate. Since this fitting formulae is still not widely used, compared to Peacock \& Dodds (1996) formulae, we reproduce them here for interested readers. The nonlinear power spectrum is related to the linear power spectrum through (Ma 1998): \begin{eqnarray} && {\Delta(k)\over \Delta_L(k_L)} = G\left({\Delta_L(k_L) \over g_0^{1.5}\,\sigma_8^\beta} \right) \,,\nonumber\\ && G(x)=[1+\ln(1+0.5\,x)]\,{1+0.02\,x^4 + c_1\,x^8/g^3 \over 1+c_2\,x^{7.5}}\,, \end{eqnarray} where $\Delta(k)\equiv k^3P_\delta(k)/(2\pi^2)$ is the density variance in the linear and nonlinear regimes\footnote{The notations used by Peacock \& Dodds (1996) and Ma (1998) differs in that $P(k)$ defined in Ma (1998) refers to $P(k)/(2\pi)^3$ in Peacock \& Dodds (1996).}. Similar to Peacock \& Dodds (1996), the nonlinear scale is related to the linear scale through: \begin{equation} k_L =\left[1+ \Delta(k)\right]^{-1/3}k. \end{equation} Instead of the effective spectral index $n_{\rm eff}$ used in Peacock \& Dodds (1996), the formalism uses $\sigma_8$ which is the rms linear mass fluctuation on $8\,h^{-1}$ Mpc scale evaluated at the redshift of interest. The numerical simulations suggest that $n_{\rm eff}+3$ is related to $\sigma_8$ through $n_{\rm eff}+3 \sim \sigma_8^\beta$ where $\beta=0.7+10\Omega_\nu^2$. The functions $g_0=g(\Omega_m,\Omega_\Lambda)$ and $g=g(\Omega_m(z),\Omega_\Lambda(z))$ are, respectively, the relative growth factor for the linear density field evaluated at present and at redshift $z$, for a model with a present-day matter density $\Omega_m$ and a cosmological constant $\Omega_\Lambda$. A fitting formula for $g(\Omega_m,\Omega_\Lambda)$ is (Carroll et al. 1992): \begin{eqnarray} && g ={5\over 2}\Omega_m(z) \, \\ && [ \Omega_m(z)^{4/7}-\Omega_\Lambda(z)+\left(1+ \Omega_m(z)/2\right) \left(1+ \Omega_\Lambda(z)/70 \right)]^{-1}\, \nonumber . \end{eqnarray} According to Ma (1998), for CDM and LCDM models, a good fit is given by $c_1=1.08\times 10^{-4}$ and $c_2=2.10\times 10^{-5}$, while $c_1=3.16\times 10^{-3}$ and $c_2=3.49\times 10^{-4}$ for MDM models with $\Omega_\nu$ of $\sim$ 0.1 and $c_1 =6.96\times 10^{-3}$ and $c_2=4.39\times 10^{-4}$ for MDM models with $\Omega_\nu$ of $\sim$ 0.2. For all other $\Omega_\nu$ values, usually less than 0.1 for neutrino masses of current interest, we interpolate between the published values of $c_1$ and $c_2$ by Ma (1998). This procedure should be approximate, but for higher precision, numerical simulations would be required to determine $c_1$ and $c_2$ at individual $\Omega_\nu$ values. In general, the weak lensing convergence power spectrum depends on six cosmological parameters: $\Omega_m$, $\Omega_\Lambda$, $\Omega_b$, $\Omega_\nu$, $n_s$ the primordial scalar tilt and $\delta_H$ the normalization of the density power spectrum. Also, throughout this paper, we take a flat model in which $\Omega_m+\Omega_\Lambda=1$. Such a cosmology is motivated by both inflationary scenarios and current observational data. For $\delta_H$ we use COBE normalizations as presented by Bunn \& White (1997) and also consider galaxy cluster based normalizations, $\sigma_8$, from Viana \& Liddle (1998). \begin{figure*} \centerline{\psfig{file=fig1.eps,width=5.0in,angle=-90}} \caption{Weak Lensing Power Spectrum for two COBE normalized models involving non-zero mass neutrinos. The upper set of curves correspond to a flat cosmological model involving $\Omega_m=0.75$, $\Omega_b=0.05$, $h=0.65$, $n_s=1$, while the lower set of curves are for $\Omega_m=0.35$ with other parameters as above. In {\it grey}, we show weak lensing power spectra using Peacock \& Dodds (1996) fitting function for nonlinear evolution of the power spectrum, calculated assuming its validity for MDM cosmologies, while dark lines show weak lensing power spectra for recently updated fitting functions for MDM cosmologies by Ma (1998). In {\it dot-dashed} lines, we show expected errors in the power spectrum measurement from a weak lensing survey of 25 $\times$ 25 deg$^{2}$ down to the magnitude limits of 25 in R. Such a survey is expected to be available in near future with wide-field cameras such as MEGACAM (Boulade et al. 1998).} \end{figure*} In Fig.~1, we show two sets of COBE normalized weak lensing power spectra considering the presence of non-zero mass neutrinos. The upper and lower curves represent two cosmological models with high and low $\Omega_m$ values and computed assuming the redshift of background sources as given in Eq.~4, with $\beta=1.5$ and $\alpha=2.0$. As shown, non-zero mass neutrinos suppress power at large $l$ values, and this effect is significant for low $\Omega_m$ models. This is primarily due to that fact that the suppression of power is directly proportional to the ratio of $\Omega_\nu/\Omega_m$. In addition, we have also shown the expected 1$\sigma$ uncertainty in the power spectrum measurement for a survey of size 625 deg$^{2}$ down to magnitude limit of 25 in R. It is likely that weak lensing surveys down to R of 25 within an area of 100 deg$^2$ will be available in the near future, and that the area coverage would steadily grow as high as several thousand square degrees over the next decade. As shown in Fig.~1, reliable measurements of the power spectrum is likely when $l$ is between 100 and 3000. This is the same range in which neutrinos suppress power. Such effects do not exist, for example, in the CMB anisotropy power spectrum; low redshift probes of the matter power spectrum provide ideal ways to weigh neutrinos. \subsection{Cosmic Confusion?} However, there are alternative possibilities which can mimic neutrinos. In Fig.~2, as examples, we illustrate two possibilities which can produce a similar power spectrum as a model involving $\Omega_m$ of 0.35 and $m_\nu$ of 0.7 eV; When $m_\nu$ is 1 eV, increasing the primordial scalar tilt by 30\% can mimic the original power spectrum, while in a model with zero mass neutrinos, increasing the baryon content by 80\% can produce essentially the original power spectrum. Such effects are essentially what can be described as {\it cosmic confusion}, and thus, careful measurements of cosmological parameters are needed to weigh neutrinos even with weak lensing. \begin{figure} \centerline{\psfig{file=fig2.eps,width=3.6in,angle=-90}} \caption{Cosmic confusion: Two alternate models involving changes in the baryon content or the scalar tilt produce essentially the same power spectrum as a model involving 0.7 eV neutrinos. The {\it grey} curves are same as Fig.~1.} \end{figure} \section{Neutrino Mass Measurement} In order to investigate the possibility for a neutrino mass measurement, we consider the so-called Fisher information matrix (e.g., Tegmark et al. 1997) with six cosmological parameters that define the weak lensing power spectrum. The Fisher matrix $F$ can be written as: \begin{equation} F_{ij} = -\left< \partial^2 \ln L \over \partial p_i \partial p_j \right>_{\bf x}, \end{equation} where $L$ is the likelihood of observing data set ${\bf x}$ given the parameters $p_1 \ldots p_n$. Following the Cram\'er-Rao inequality, no unbiased method can measure the {\it i}th parameter with standard deviation less than $(F_{ii})^{-1/2}$ if other parameters are known, and less than $[(F^{-1})_{ii}]^{1/2}$ if other parameters are estimated from the data as well. Since Eq.~6 is usually calculated assuming a prior cosmological model, the estimated errors on the parameters of this underlying model can be dependent on prior assumptions. Assuming a Gaussian and uncorrelated distribution for uncertainties, one can easily derive the Fisher matrix for weak lensing as\footnote {Note the minor correction to Eq.~4 of Hu \& Tegmark (1999)}: \begin{equation} F_{ij} = \sum_{l=l_{\rm min}}^{l_{\rm max}} \frac{f_{\rm sky} (2l +1)}{2 \left(P_\kappa(l)+\frac{\langle \gamma^2 \rangle}{n_{\rm mag}}\right)^2} \frac{\partial P_\kappa(l)}{\partial p_i} \frac{\partial P_\kappa(l)}{\partial p_j}. \end{equation} As illustrated in Fig.~2, in order to make a reliable measurement of neutrino mass, it is necessary that one consider external measurements of cosmological parameters. Such measurements can come from variety of probes such as Type Ia supernovae, galaxy clusters, CMB, gravitational lensing etc. Here, we take both a conservative approach with large uncertainties for the cosmological parameters based on other techniques and a more optimistic approach motivated by the expected uncertainties from future surveys. In our conservative model, we use following errors: $\sigma(\Omega_m)=0.2$, $\sigma(\Omega_b) = 0.1 \Omega_b$, $\sigma(h) =0.2$, $\sigma(n_s)=0.1$, $\sigma(\ln \delta_H) = 0.5$, while in our optimistic model, we use $\sigma(\Omega_m)=0.07$, $\sigma(\Omega_b) = 0.0025h^{-2}$, $\sigma(h) =0.1$, $\sigma(n_s)=0.06$, $\sigma(\ln \delta_H) = 0.3$. These errors are in fact worser than what is expected to be measured from PLANCK\footnote{http://astro.estec.esa.nl/Planck/; also, ESA document D/SCI(96)3.}, but is similar to what could be achieved with a mission such as MAP\footnote{http://map.gsfc.nasa.gov/}. We consider a fiducial model in which $\Omega_b=0.019$, consistent with current observations, $h=0.65$ and $n_s=1.0$ and normalization based on COBE. In addition, we also consider an alternative normalization to the power spectrum based on measurements of $\sigma_8$ ($=0.56\Omega_m^{-0.47}$) following Viana \& Liddle (1998). We also consider variations to the above fudicial model and marginalize over the uncertainties to obtain the 2$\sigma$ detection limit of neutrinos for various weak lensing surveys. We only use the information on the power spectrum between $l$ values of 100 and 5000. At $l$ values below 100, cosmic variance dominate the measurement while at $l > 5000$ the finite number of galaxies and their ellipticies contribute to the increase in power spectrum measurement uncertainties. In Fig.~3, we summarize our results: solid lines show the expected 2$\sigma$ detection limit for our conservative errors while dashed lines show the detection limits for more optimistic errors. The dot-dashed line is for models in which the matter power spectrum is normalized to 8 h$^{-1}$ Mpc scales. The high dependence of its value and error on $\Omega_m$ causes the $\sigma_8$ normalized limits to be different from those in which power spectra are normalized to COBE measurements. In Fig.~3, we have shown the limits assuming a survey of 100 $\times$ 100 sqr. degrees down to a R band magnitude of 25. However, for surveys with different areas, especially for surveys in near future with small coverage, the limits can be scaled by the reduction factor in the observed area (see, Eq.~12). We assume uncorrelated errors in the weak lensing power spectrum measurement. For low $\Omega_m$ models ($\la 0.5$) normalized to COBE, and using our conservative errors, we can write the 2 $\sigma$ detection limit on the neutrino mass as: \begin{equation} m_\nu^{2\sigma} \sim 5.5 \left(\frac{\Omega_mh^2}{0.1 N}\right)^{0.8} \left(\frac{10}{\theta_s}\right). \end{equation} This 2 $\sigma$ detection limit is comparable to current upper limits at the 2 $\sigma$ level on the neutrino mass. Using more optimistic errors decreases this limit by a factor of 2 to 3 depending on $\Omega_m$, however, to obtain such optimistic errors on cosmological parameters one require accurate measurements on the CMB power spectrum such as to the level of MAP satellite. In making this prediction we have assumed that the weak lensing power spectrum can be measured to the expected uncertainty predicted by simple arguments involving errors in ellipticities and cosmic confusion and that the measurements are uncorrelated. Also, in order to obtain a reliable measurement of the weak lensing power spectrum, one require additional knowledge on the redshift distribution of sources. Such information is likely to be adequately obtained with photometric redshift measurements of color data or by template fitting techniques that has been developed for multicolor surveys (e.g., Hogg et al. 1998). The accuracy to which such measurements can be made should be adequate, however, if no multicolor data is available then this may not be possible. Therefore, it is likely that such a {\it clean} measurement of the weak lensing power spectrum will not be directly possible in the near future. In order to consider such affects, we increased the expected uncertainties in the power spectrum by a factor of 2 beyond what is predicted for a survey of 100 sqr. degrees down to a R band magnitude of 25. The expected neutrino mass limit increases by an amount consistent with what is expected from the Fisher matrix formalism. Even in such a scenario with a poorly measured power spectrum, one can still put interesting limits on the neutrino mass. A small area survey such as 10 $\times$ 10 sqr. degrees is likely to be feasible in the near future with upcoming observations from wide field CCD cameras. There are several such instruments currently either in the design or manufacturing stages: MEGACAM\footnote{ http://cdsweb.u-strasbg.fr:2001/projects/megacam/} which will make observations from the Canada France Hawaii Telescope (CFHT; Boulade et al. 1998), VLT-Survey-Telescope\footnote{ http://oacosf.na.astro.it/vst/} (VST). Other than these surveys, which are likely to first produce deep weak lensing surveys over small areas, two wide-field shallow surveys are currently ongoing at optical (SDSS; Stebbins et al. 1997) and radio (FIRST; Kamionkowski et al. 1997), however, it is still unclear as to what accuracy these imaging data can be used for weak lensing studies. Still, assuming that SDSS can in fact make weak lensing measurements down to a R band magnitude of 22, we find that given its wide field coverage, it can also be used to detect neutrinos down to a mass limit of $\sim$ 3 eV at the 2$\sigma$ level, or to put interesting limits at the same mass threshold. For an ultimate survey of $100 \times 100$ deg$^2$, weak lensing allows a detection of neutrinos down to a mass of $\sim$ 0.5 eV when $\Omega_m \sim 0.3$ and $h \sim 0.65$. With expected errors from CMB satellites, this limit can be lowered by a factor of 3 to 4 allowing a possibility for weak lensing surveys to probe neutrinos with mass lower than 0.1 eV. These conclusions, generally, are consistent with what was found by Hu \& Tegmark (1999); minor differences are likely to arise from the fact that the present study and Hu \& Tegmark (1999) used different fitting functions to describe the non linear evolution of the potential power spectrum and that fudicial cosmological models may be different. We note here that using MAP or PLANCK data with galaxy redshift surveys such as from SDSS, and no weak lensing measurements, only allow the determination of neutrino mass to a limit of $\sim$ 1 eV and 0.3 eV respectively (Hu et al. 1997). \begin{figure} \centerline{\psfig{file=fig3.eps,width=4.1in,angle=90}} \caption{Expected 2$\sigma$ detection limit for 100 $\times$ 100 deg.$^2$ weak lensing survey down to a magnitude limit of 25, assuming a spatially flat Universe with a scalar tilt of 1. Solid line represents detection with our conservative errors while the dashed line represent detection with more optimistic errors. The dot-dashed line is for models involving normalizations based on current measurements of $\sigma_8$ and using more optimistic errors.} \end{figure} Returning to much smaller surveys, we have only studied the accuracy to which the neutrino mass can be measured. However, in making such measurements one does not lose information to make other measurements as well. For example, the conservative errors we assumed on other cosmological parameters can also be improved by factors of 2 to 3 when information on these parameters are also derived with weak lensing. Also, one can abandon the assumption of a spatially flat Universe, and determine the value of the cosmological constant directly from weak lensing data, while also putting a limit on the neutrino mass. However, if the assumption on a spatially flat Universe is dropped in order to measure $\Omega_\Lambda$, then the limit to which neutrino mass can be measured increases by a factor of $\sim$ 1.5 for surveys of size 100 sqr. degrees. For now, if one to measure or improve all other cosmological parameters that can be studied with weak lensing surveys (and listed in Sect.~2.1), then it is safe to say that neutrinos down to a mass limit of $\sim$ 8 eV can be measured with weak lensing surveys of size 100 sqr. degrees down to a R band magnitude of 25. Such a possibility will definitely be available with upcoming surveys from MEGACAM. For still smaller surveys, such as 10 sqr. degrees, if one attempts to make all cosmological measurements, such as $\Omega_m$ and $\Omega_\Lambda$ interesting limits on neutrino mass can only be obtained at a mass level greater than 25 eV. Since such neutrino masses may be ruled out, it is safe to ignore the presence of neutrinos when making measurements with much smaller surveys. Such surveys are likely to be first available with wide-field cameras, with the coverage increasing afterwards. \section{Discussion \& Summary} Here, we have considered the possibility for a neutrino mass measurement using weak gravitational lensing of background sources due to foreground large scale structure. For survey of size 100 deg$^2$, neutrinos with masses greater than $\sim$ 5.5 eV could easily be detected. This detection limit is comparable to the current cosmological limits on neutrino mass, such as from the Ly$\alpha$ forest. When compared to various ongoing experiments to detect neutrinos, the advantage of weak lensing is that one can directly obtain a measure of mass rather than mass difference between two neutrino species. For typical surveys of size $\sim$ ten square degrees, ignoring the presence of neutrinos can lead to biased estimates for cosmological parameters, e.g., cosmological mass density can be underestimated by a factor as high as $\sim$ 15\% if neutrinos with mass 5 eV are in fact present. However, if such weak lensing surveys are solely used for the derivation of parameters such as cosmological mass density, than the accuracy to which such derivations can be made is less than the bias produced by neutrinos. Therefore, for small area surveys, the presence of neutrinos can be safely ignored (assuming that their masses is less than $\sim$ 5 eV or so). However, armed with cosmological parameters from other complimentary techniques, even such small weak lensing surveys allow a strong possibility to investigate the presence of non-zero mass neutrinos. \begin{acknowledgements} We acknowledge useful discussions with Wayne Hu and Dragan Huterer. Wayne Hu is also thanked for communicating the fitting code to evaluate the MDM transfer function. We also thank an anonymous referee for comments which led to several improvements in the presentation. \end{acknowledgements}
\section{Introduction} Massive black hole s (MBHs) have been postulated in quasars and active galaxies (Lynden-Bell 1969, Rees 1984). Evidence for the existence of MBHs has recently been found in the center of our Galaxy (Ghez et al.\ 1998, Genzel et al.\ 1997) and in the weakly active galaxy NGC 4258 (Miyoshi et al.\ 1995). Compact dark masses, probably MBHs, have been detected in the cores of many normal galaxies using stellar dynamics (Kormendy and Richstone 1995). The MBH mass appears to correlate with the galactic bulge luminosity, with the MBH being about one percent of the mass of the spheroidal bulge (Magorrian et al.\ 1998, Richstone et al.\ 1998). The question whether AGN follow a similar black hole -bulge relation as normal galaxies is a very interesting one, as it may shed light on the connection between the host galaxy and the active nucleus. Wandel \& Mushotzky (1986) have found an excellent correlation between the virial mass included within the narrow line region (of order of tens to hundreds pc from the center) and the black hole~mass estimated from X-ray variability in a sample of Seyfert 1 galaxies. A black hole -bulge relation similar to that of normal galaxies has been reported between MBH of bright quasars and the bulge of their host galaxies (Laor 1998), but the black hole~ and bulge mass estimates have large uncertainties (section 3.2). Seyfert 1 galaxies provide an opportunity to obtain more reliable black hole -to- bulge mass ratios (BBRs): because of their lower nuclear brightness, their bulge magnitudes can be measured directly. Also the black hole~ mass estimates (Wandel 1998) are much more reliable for AGN with reverberation data, which are more readily obtained for low luminosity AGN. The relation between the bulge and the nonstellar central source has been studied for many Seyfert galaxies (Whittle 1992; Nelson \& Whittle 1996). These works find a tight correlation between the stellar velocity dispersion and the O[III] line and radio luminosity. Reliable BLR size measurements are now possible through reverberation mapping techniques (Blandford \& McKee 1982, recently reviewed by Netzer \& Peterson 1997). High quality reverberation data and virial masses are presently available for about twenty AGN, most of them Seyefert 1s (Wandel, Peterson and Malkan 1999, hereafter WPM). We combine the reverberation~ masses (section 2) with Whittle's bulge estimates in order to study the BBR in low-luminosity AGN and compare it to MBHs in normal galaxies and quasars (section 3). In section 4 we derive a MBH-evolution theory that can explain our results. \section{BLR reverberation as a probe of black hole masses in AGN} Broad emission lines probably provide the best probe of black hole s in AGN. Assuming the line-emitting matter is gravitationally bound, and hence has a near-Keplerian velocity dispersion (indicated by the line width), it is possible to estimate the virial central mass: $M\approx G^{-1}rv^2 .$ This remains true for many models where the line emitting gas is not in Keplerian motion, such as radiation-driven motions and disk-wind models (e.g. Murray et al.\ 1998): in a diverging outflow the density (end hence the emissivity) decreases outwards, the emission is dominated by the gas close to the base of the flow, where the velocity is close to the escape velocity. (Note that if the velocity is actually larger than Keplerian, the virial mass is an upper limit and the result that the Seyfert galaxies in our sample have smaller black hole~ masses than MBHs detected in normal galaxies becomes even stronger). The main problem in estimating the virial mass from the emission-line data is to obtain a reliable estimate of the size of the BLR, and to correctly identify the line width with the velocity dispersion in the gas. WPM use the continuum/emission-line cross-correlation function to measure the responsivity-weighted radius $c\tau$ of the BLR (Koratkar \& Gaskell 1991), and the variable (rms) component of the spectrum to measure the velocity dispersion in the same part of the gas which is used to calculate the BLR size, automatically excluding constant features such as narrow emission lines and Galactic absorption. The line width and the BLR size yield the virial "reverberation" mass estimate $M_{rev} \approx (1.45\times 10^5M_{\odot} ) c\tau_{days}v_3^2$ where $v_3$ is the rms FWHM in units of $10^3 \ifmmode {\rm km\ s}^{-1} \else km s$^{-1}$\fi$. The virial assumption ($v \propto r^{-1/2}$) has been directly tested using data for NGC 5548 (Krolik et al.\ 1991; Peterson \& Wandel 1999). The latter authors find that when the BLR reverberation size is combined with the rms line width in multi-year data for NGC 5548, the virial masses derived from different emission lines and epochs are all consistent with a single value ($(6.3\pm 2)\times 10^7M_{\odot}$) which demonstrates the case for a Keplerian velocity dispersion in the line-width/time-delay data. \section {The Black-Hole - Bulge Relation } \subsection{Seyfert 1 galaxies} We use the WPM sample with the virial mass derived from the H$\beta$ line by the reverberation-rms method (table 1). For 13 of the objects in the WPM sample we obtain the bulge magnitudes from the compilation of Whittle et al.\ (1992), who calculate the bulge magnitude from the total blue magnitude, using the empirical formula of Simien \& deVaucolours (1986), relating the galaxy type to the bulge/total fraction. The bulge magnitudes are corrected for the nonstellar emission using the correlation between H$\beta$ and the nonstellar continuum luminosity (Shuder 1981). Mkn 110 and Mkn 335 have no estimated bulge magnitude in Whittle's compilation, because they do not have well defined Hubble types. For these objects we adopt a canonical Hubble type of Sa, which has a bulge correction ($m_{bulge}-m_{gal}$) of 1.02 mag. For Mkn 335 there is already a fairly large (and therefore uncertain) correction for the active nucleus (1.17 mags). For 3C120 the bulge magnitude is taken from Nelson \& Whittle (1995) who find a bulge magnitude of -22.12. The uncertainties in the bulge magnitude were estimated from Whittle's (1992) quality indicators. These indicators estimate the error in the subtraction of the nonstellar luminosity and some other factors, which for most galaxies amount to an uncertainty in the range 0.2-0.6 magnitudes. To the galaxies with an uncertain Hubble type we assign an uncertainty of 1.2 mag. \begin{table} \caption{AGN Central Masses derived from BLR data, Compared with Host Bulge magnitudes and corresponding mass. Column (2) -- FWHM of (H$\beta$), rms profile, in $10^3$km/s. (3) - log(lag) -- corresponding to the BLR size light days, (4) -- absolute blue bulge magnitude from Whittle et al.\ (1992), (5) -- log of the galactic bulge mass ($M_{bul}$) in $M_{\odot}$, (6) -- black hole mass (from WPM), (7) -- BH to bulge mass ratio, (8) -- the Eddinton ratio of the ionizing luminosity. \label{tbl-2}} \begin{center \begin{tabular}{llllllll} \tableline {}&{}&{}&{}&{}&{}&{}&{}\\ Name & FWHM & log($\tau$)& $-M_B$ & log$(M_{bul})$ & ${M_{bh}\over 10^7M_{\odot}}$ &log(${M_{bh}\over M_{bul}}$)& log(${L_{ion}\over L_E}$)\\ {(1)}&{(2)}&{(3)}&{(4)}&{(5)}&{(6)}&{(7)}&(8)\\ \tableline {}&{}&{}&{}&{}&{}&{}&{}\\ 3C\,120\tablenotemark{a} &$ 2.2$& 1.64& 20.3$\pm$ 1.2 & 10.77$\pm$ 0.5 & $ 3.1^{+2.0}_{-1.5}$ &-3.28&-0.57\\ 3C\,390.3 &$ 10.5$& 1.38& 22.12$\pm$0.4 & 11.0$\pm$ 0.15 & $39.1^{+12.4}_{-14.8}$ &-2.41&-2.19\\ Akn\,120 &$ 5.85$& 1.59& 21.06$\pm$0.8& 11.12$\pm$0.3 & $19.4^{+4.1}_{-4.6}$&-2.83 &-1.48\\ F\,9 &$ 5.9$& 1.23& 22.25$\pm$0.6& 11.67$\pm$0.25 & $ 8.7^{+2.6}_{-4.5}$ &-3.73&-1.84\\ IC\,4329A &$ 5.96$& 0.15& 19.93$\pm$0.8& 10.60$\pm$0.3 & $<0.73$ &$<-3.73$ &$>-2.94$\\ Mrk\,79 &$ 6.28$& 1.26& 20.19$\pm$0.2& 10.72$\pm$0.1 & $10.4^{+4.0}_{-5.7}$ &-2.70&-1.87\\ Mrk\,110\tablenotemark{a} &$ 1.67$& 1.29& 20.76$\pm$1.0& 10.98$\pm$0.4 & $0.80^{+0.29}_{-0.30}$ &-4.07&-0.69 \\ Mrk\,335\tablenotemark{a} &$ 1.26$& 1.23& 20.02$\pm$1.0 & 10.64$\pm$0.4& $0.39^{+0.14}_{-0.11}$ &-4.06&-0.49\\ Mrk\,509 &$ 2.86$& 1.90& 21.75 $\pm$0.6& 11.44$\pm$0.25 & $9.5^{+1.1}_{-1.1}$ &-3.46&-0.54\\ Mrk\,590 &$ 2.17$& 1.31& 21.26$\pm$0.2 & 11.21$\pm$0.1 & $1.4^{+0.3}_{-0.3}$ &-4.06&-0.89 \\ Mrk\,817 &$ 4.01$& 1.19& 21.17$\pm$0.4 & 11.17$\pm$0.15 & $ 3.7^{+1.1}_{-0.9}$ &-3.61&-1.54\\ NGC\,3227 &$ 5.53$& 1.04& 20.46$\pm$0.4 & 10.84$\pm$0.25 & $4.9^{+2.7}_{-4.9}$ &-3.15 &-1.98\\ NGC\,3783 &$ 4.1$& 0.65& 20.07$\pm$0.2 & 10.66$\pm$0.1 & $1.1^{+1.1}_{-1.0}$ &-3.62&-2.10\\ NGC\,4051 &$ 1.23$& 0.81& 19.70$\pm$0.2 & 10.62$\pm$0.1 & $0.14^{+ 0.15}_{-0.09}$ &-4.42&-0.86\\ NGC\,4151 &$ 5.23$& 0.48& 19.98$\pm$0.4 & 11.04$\pm$0.15 & $1.2^{+ 0.8}_{-0.7}$ &-3.54 &-2.49\\ NGC\,5548 &$ 5.50$& 1.26& 20.89$\pm$0.2& 11.05$\pm$0.1 & $6.8^{+ 1.5}_{-1.0}$ &-3.06&-1.83\\ NGC\,7469 &$ 3.2$& 0.70&20.90$\pm$0.2& 11.05$\pm$0.1 & $0.76^{+ 0 .75}_{-0.76}$& -4.15& -1.85\\ PG\,0953+414&$ 3.14$& 2.03& 20.29$\pm$1.0\tablenotemark{b}& 11.49$\pm$0.4 & $15.5^{+ 10.8}_{-9.1}$ &-2.57&-0.49\\ {}&{}&{}&{}&{}&{}&{}&{}\\ \tableline \end{tabular} \tablenotetext{a}{Unknown Hubble type, bulge correction estimated assuming Sa} \tablenotetext{b}{From Bahcall et al.\ (1997)} \end{center} \end{table} We relate the bulge luminosity to the magnitude by the standard expression ${\rm log} (L_{bulge}/L_\odot )= 0.4 (-M_v + 4.83)$. The bulge mass is then calculated using the mass-to-light relation for normal galaxies, ${M/M_{\odot} \over L/L_\odot}~\approx 5 (L/10^{10}L_\odot )^{0.15}$ (see Faber et al.\ 1997). Fig. 1 shows the black hole~ mass as a function of the bulge mass. All the objects in our sample have BBRs lower than 0.006, the average value found for normal galaxies (Magorrian et al.\ 1998, represented by a dashed line), and the sample average is $<M_{BH}>= 3\times 10^{-4} <M_{bulge}>$. Also shown is NGC 1068 (a Seyfert 2), with the MBH mass estimated by maser dynamics. The narrow-line Seyfert galaxy NGC 4051, which has by far the lowest BBR in our sample, may indicate that narrow-line Seyfert 1 galaxies have smaller black hole s than ordinary Seyfert 1 galaxies (Wandel and Boller 1998). \myfig {12} 1 {fig1pc.eps} {1. The virial black hole ~mass calculated by the reverberation BLR method (from Wandel, Peterson \& Malkan 1999) vs. the bulge magnitude (from Whittle 1992) for the Seyfert 1 galaxies in our sample (diamond), the masing Seyfert 2 galaxy NGC 1068 (square) and PG0953+414 (triangle). Open diamonds indicate an unknown Hubble type (and therefore a large uncertainty in the bulge magnitude). The dashed diagonal lines are the average BBRs for normal galaxies (Magorrian et al.\ 1998) and Seyfert 1s (this work). } \subsection {Quasars} Laor (1998) has studied the black hole -host bulge relation for a sample of 15 bright PG quasars. Estimating the bulge masses from the Bahcall et al.\ (1997) study of quasar host galaxies he admits the uncertainty in estimating bulge luminosity, dominated by the much brighter nonstellar source. Laor estimates the black hole~ mass using the H$\beta$ line width and the empirical relation $r_{BLR}=15 L_{44}^{1/2}~ ~{\rm light-days}$ (Kaspi et al.\ 1997), where $L_{44}=L(0.1-1\micron )$ in units of $10^{44}\ifmmode {\rm erg\ s}^{-1} \else erg s$^{-1}$\fi$. As this relation has been derived for less than a dozen low- and medium luminosity objects (mainly Seyferts) with measured reverberation sizes, it is not obvious that it may be extrapolated to more luminous quasars. The BLR size is also dependent on the ionizing and soft X-ray continua (Wandel 1997). The WPM sample (which includes Kaspi's sample) indicates that the slope of the BLR-size luminosity relation may flatter than 0.5; WPM find $r\sim 17L_{44}^{0.36\pm 0.09}{\rm l-d}$. If this result is correct, extrapolating the $r\sim L^{1/2}$ relation over two orders of magnitude (the difference between the average luminosity of the PG quasars used by Laor and Kaspi's sample average) overestimates the black hole~ mass. Indeed, for the only object common to the Laor and WPM samples - the quasar PG 0953+414 - Laor finds 3$\times 10^8M_{\odot}$, while the reverberation~ -rms method gives $(1.5^{+1.1}_{-0.9})\times 10^8 M_{\odot}$. \subsection {Comparing Normal Galaxies, Seyferts and Quasars} Fig. 2 shows the three groups in the plane of black hole~mass vs. bulge luminosity. The best fits and the corresponding standard deviations to the data in the three groups are ($M_8=M_{BH}/10^8M_{\odot}$ and $L_{10}=L_{bulge}/10^{10}L_\odot$): \begin{enumerate} \item Normal galaxies (Magorrian et al.\ 1998, table 2, excluding upper limits) - $M_8= 2.9 L_{10}^{1.26}$, $\sigma = 0.47$ \item PG quasars (Laor 1998, all objects in his table 1) - $M_8= 1.6 L_{10}^{1.10}$, $\sigma = 0.38$ \item Seyfert 1s (this work, excluding NGC 4051) $M_8= 0.2 L_{10}^{0.83}$, $\sigma = 0.43$ \end{enumerate} \myfig {13} {1} {fig2pc.eps} {2. Mass estimates of MBHs plotted against the luminosity of the bulge of the host galaxy. Squares: MBH candidates from Magorrian et al.\ (1998), open squares - MBHs detected by maser dynamics triangles - PG quasars from Laor (1998), diamonds - Seyfert 1 galaxies (this work). MW denotes our Galaxy. Also given are the best linear fits for each class (see text). The dashed long line is the estimate of dead black hole s from integrated AGN light. } As a group Seyfert 1 galaxies have a significantly lower BBR than normal galaxies and bright quasars. This lower value agrees with the remnant black hole~ density derived from integrating the emission from quasars (Chokshi and Turner 1992): $\rho_{BH}=\int\int (L/\epsilon c^2)\Phi(L,t)dLdt =2\times 10^5(\epsilon /0.1) M_{\odot} {\rm Mpc}^{-1}$, ($\Phi$ is the quasar luminosity function and $\epsilon$ is the efficiency), which compared to the density of starlight in galaxies ( $\rho_{gl}$) gives $\rho_{BH}/\rho_{gl}= 2\times 10^{-3} (0.1/\epsilon )(M_{\odot} /L_\odot )$ (shown as a dashed line in Fig. 2). \section {Black hole Evolution and the Black Hole - Bulge ratio} \subsection{Demography} While Seyfert 1 galaxies seem to have a lower BBR than bright quasars and the galaxies with {\it detected} MBHs in the Magorrian et al.\ (1998) sample, they are in good agreement with the BBRs of the {\it upper limits} and of our Galaxy, and with remnant quasar black~holes. It is plausible therefore that the Seyfert galaxies in our sample represent a larger population of galaxies with low BBRs, which is under-represented in the Magorrian et al.\ sample. This hypothesis is supported by the distribution of black hole~ masses in Fig. 2: the only MBHs under ~$2\times 10^8M_{\odot}$ detected by stellar dynamical methods are in the Milky Way, in Andromeda and its satellite M32, and in NGC 3377 (the latter being nearly $10^8M_{\odot} $). These galaxies, as well as NGC1068 and at least two of the three upper limit in Magorrian's sample do have low BBRs, comparable to our Seyfert 1 average. Actually for angular-resolution limited methods, the MBH detection limit is correlated with bulge luminosity: for more luminous bulges the detection limit is higher, because the stellar velocity dispersion is higher (the Faber-Jackson relation). In order to detect the dynamic effect of a MBH it is necessary to observe closer to the center, while the most luminous galaxies tend to be at larger distances, so for a given angular resolution, the MBH detection limit is higher. This may imply that Magorrian et al.\ `s sample is biased towards larger MBHs, as present stellar-dynamical methods are ineffective for detecting MBHs below $\sim 10^8M_{\odot}$ (except in the nearest galaxies). The BLR method is not subject to this constraint, making Seyfert 1 galaxies good candidates for detecting low-mass MBHs. (Note however that by the same token the WPM sample may be biased towards Seyferts with low black hole~masses, which tend to vary on shorter timescales and hence are more likely to be chosen for reverberation studies). \subsection{Black Hole Growth by Accretion} Fig. 2 shows that Seyfert 1 galaxies have relatively small MBHs compared MBHs in normal galaxies and to quasars, yet they have comparable bulges. Below we suggest a possible explanation. Consider MBH growth by accretion from the host galaxy. Since the accretion radius, $R_{acc}\approx 0.3 M_6 v_2^{-2} $pc (where $M_6=M_{BH}/10^6M_{\odot} $ and $v_2=v_*/100\ifmmode {\rm km\ s}^{-1} \else km s$^{-1}$\fi$ is the stellar velocity dispersion) is small compared with the size of the bulge, we may assume the mass supply to the black hole~is given by the spherical accretion rate, $\dot M= 4\pi\lambda R_{acc}{v_*}^2 \rho = (10^{-4}M_{\odot} /{\rm yr}) \lambda M_6^2v_2^{-3}\rho_*$, where $\rho_*$ is the stellar (or gas) mass density in units of $ M_{\odot} pc^{-3}$ (corresponding to $4.4 g/cm^3$) and $\lambda<1$ is the Bondi parameter combined with a possible reduction factor due to angular momentum. Integrating we find the time required for growing from a mass $M_i$ to $M_f$ by accretion of gas or stars, $t_{acc}= (10^{16} yr) v_2^3\rho_*^{-1}\lambda^{-1} ( M_{\odot} / {M_i} - M_{\odot} / {M_f} ) \approx (10^8 {\rm yr})M_8^{-1}v_2^3\rho_*^{-1}$. For masses $<10^6\rho_*^{-1}M_{\odot}$ this is larger than the Hubble time, so seed MBHs must grow by black hole~ coalescence which, even for dense clusters, is of the order of the Hubble time (Lee, 1993; Quinlan \& Shapiro 1987). For densities as high as in the central parsec of the Milky Way (few$\times 10^7 M_{\odot} pc^-3$; Genzel et al.\ 1997) or for NGC 4256 (Miyoshi et al.\ 1995) accretion-dominated growth becomes feasible for masses as low as 100-1000$M_{\odot}$. While the accretion rate is growing as $M^2$ and the growth time decreases as $M^{-1}$, the MBH eventually becomes large enough for accretion-dominated growth time $t_g=t_{acc}$. This phase may be applicable for the Seyfert population. Since the luminosity is $L\propto\dot M\propto M^2$, the Eddington ratio increases as $L/L_{Edd}\propto M\propto L^{1/2}$. The black hole~growth slows down when the Eddington ratio approaches unity, $t_g$ being bound by the Eddington time, $ t_g\sim t_E=M/ {\dot M_E} = 4.5\times 10^7 ( \epsilon /{0.1} )^{-1} {\rm yr} $ where $\dot M_E$ is the accretion rate that would produce an Eddington luminosity. Equating $t_E$ to $t_{acc}$ we find that the growth rate flattens at a BH mass of $M_{t}\approx (2\times 10^8M_{\odot} ) v_2^{3}\rho_*^{-1}(\epsilon/0.1\lambda)$. In the Eddington~ -limited era, which may correspond to quasars, the growth rate is exponential, depleeting the available matter in the bulge on the relatively short time scale $t_{Edd}$. This leads to an asymptotic BBR, which is likely to be similar for luminous quasars and their largest remnant MBHs in normal galaxies. This scenario predicts that on average quasars should have higher Eddington~ ratios (near unity) than Seyferts, and larger BBRs. We can test the prediction from the data at hand. Estimating the bolometric luminosity of AGN with reverberation data from the lag (WPM), and of PG quasars from the relation $L_{bol}\approx 8 \nu L_\nu(3000\AA)$ (Laor 1998), we find a correlation between the Eddington~ ratio and the BBR, with Seyferts having Eddington~ ratios in the $10^{-3}-0.1$ range and a low BBR, and quasars with Eddington ratios close to unity and higher BBRs. From the Eddington~ ratio we can also infer the actual growth time, $t_g\approx t_E L/L_{E}$. For most objects in our sample $t_g$ is in the range $10^8-{\rm few}\times 10^9$ yr. \acknowledgments I acknowledge valuable discussions with Mark Whittle Gary Kriss, Geremy Goodman, Doug Richstone and Mark Morris and the hospitality of the Astronomy Department at UCLA.
\section{Natural Emergence of Optimized Configurations} Every day, enormous efforts are devoted to organizing the supply and demand of limited resources, so as to optimize their utility. Examples include the supply of foods and services to consumers, the scheduling of a transportation fleet, or the flow of information in communication networks within society or within a parallel computer. By contrast, {\em without\/} any intelligent organizing facility, many natural systems have evolved into amazingly complex structures that optimize the utilization of resources in surprisingly sophisticated ways (Bak 1996). For instance, driven merely by sunlight, biological evolution has developed efficient and strongly interdependent networks in which resources rarely go to waste. Even the inanimate morphology of natural landscapes exhibits patterns far from random that often seem to serve a purpose, such as the efficient drainage of water (Rodriguez-Iturbe 1997). The physical properties of these fractal patterns have aroused the interest of statistical physicists in recent times (Mandelbrot 1983). Natural systems that exhibit such self-organizing qualities often possess common features: they generally consist of a large number of strongly coupled entities with very similar properties (like species in biological evolution, despite their apparent differences). Hence, they permit a statistical description at some coarse level. An external resource (such as sunlight) drives the system which then takes its direction purely by chance. If we were to rerun evolution, there may not be trees and elephants, say, but other complex structures. Like flowing water breaking through the weakest of all barriers in its wake, species are coupled in a global comparative process that persistently washes away the least fit. In this process, unlikely but highly adapted structures surface inadvertently, as Darwin observed (Darwin 1859). Optimal adaptation thus emerges naturally, without divine intervention, from the dynamics through a selection {\em against\/} the extremely ``bad''. In fact, this process prevents the inflexibility that would inevitably arise in a controlled breeding of the ``good''. Certain models relying on extremal processes have been proposed to explain self-organizing systems in nature (Paczuski 1996). In particular, the Bak-Sneppen model of biological evolution is based on this principle (Bak 1993, Sneppen 1995). It is happily devoid of any specificity about the nature of interactions between species, yet produces salient nontrivial features of paleontological data such as broadly distributed lifetimes of species, large extinction events, and punctuated equilibrium (Gould 1977). \begin{figure} \vskip 4.40truein \special{psfile=fig1a.ps angle=-90 hoffset=-40 voffset=360 hscale=40 vscale=38} \smallskip \special{psfile=fig1b.ps angle=-90 hoffset=-40 voffset=203 hscale=40 vscale=38} \caption{Two random geometric graphs, $N=500$, with connectivities $\alpha\approx4$ (top) and $\alpha\approx8$ (bottom) in an optimized configuration found by EO. At $\alpha=4$ the graph barely percolates, with only one ``bad'' edge connecting the set of 250 round points with the set of 250 square points (diamonds show the two ends of the edge), thus $m_{\rm opt}=1$. For the denser graph on the bottom, EO obtained the cutsize $m_{\rm opt}=13$. } \label{geograph} \end{figure} Species in the Bak-Sneppen model are located on the sites of a lattice, and each is represented by a value between 0 and 1 indicating its ``fitness''. At each update step, the smallest value (representing the worst adapted species) is discarded and replaced with a new value drawn randomly from a flat distribution on $[0,1]$. Without any interactions, all the fitnesses in the system would eventually become 1. But obvious interdependencies between species provide constraints for balancing the system's overall fitness with that of its members: the change in fitness of one species impacts the fitness of an interrelated species. Therefore, at each update step in the Bak-Sneppen model, the fitness values on the sites {\em neighboring\/} the smallest value are replaced with new random numbers as well. No explicit definition is given of the mechanism by which these neighboring species are related. Yet after a certain number of updates, the system organizes itself into a highly correlated state known as self-organized criticality (SOC) (Bak 1987). In that state, almost all species have reached a fitness above a certain threshold. But these also species possess what is called punctuated equilibrium (Gould 1977): since one's weakened neighbor can undermine one's own fitness, co-evolutionary activity gives rise to chain reactions. Fluctuations that rearrange the fitness of many species occur routinely. These fluctuations can be of the scale of the system itself, making any possible configuration accessible. In the Bak-Sneppen model, the high degree of adaptation of most species is obtained by the elimination of badly adapted ones instead of a particular ``engineering'' of better ones. While such dynamics might not lead to as optimal a solution as could be engineered in specific circumstances, it provides near-optimal solutions with a high degree of latency for a rapid adaptation response to changes in the resources that drive the system. In the following we will describe an optimization method inspired by these insights (Boettcher, submitted, and Boettcher, to appear), called {\em extremal optimization\/}, and study its performance for graph partitioning and the traveling salesman problem. \section{Extremal Optimization and Graph Partitioning} In graph (bi-)partitioning, we are given a set of $N$ points, where $N$ is even, and ``edges'' connecting certain pairs of points. The problem is to partition the points into two equal subsets, each of size $N/2$, with a minimal number of edges cutting across the partition. (Call the number of these edges the ``cutsize'' $m$, and the optimal cutsize $m=m_{\rm opt}$.) The points themselves could, for instance, be associated with positions in the unit square. A ``geometric'' graph of average connectivity $\alpha$ would then be formed by connecting any two points within Euclidean distance $d$, where $N\pi d^2=\alpha$ (see Fig.~\ref{geograph}). Constraining the partitioned subsets to be of fixed (equal) size makes the solution to the problem particularly difficult. This geometric problem resembles those found in VLSI design, concerning the optimal partitioning of gates between integrated circuits (Dunlop 1985). Graph partitioning is an {\em NP-hard\/} optimization problem (Garey 1979): it is believed that for large $N$ the number of steps necessary for an algorithm to find the {\em exact\/} optimum must, in general, grow faster than any polynomial in $N$. In practice, however, the goal is usually to find near-optimal solutions quickly. Special-purpose heuristics to find approximate solutions to specific NP-hard problems abound (Alpert 1995, Johnson 1997). Alternatively, general-purpose optimization approaches based on stochastic procedures have been proposed, most notably {\em simulated annealing\/} (Kirkpatrick 1983, {\v C}erny 1985) and {\em genetic algorithms\/} (Holland 1975). These methods, although slower, are applicable to problems for which no specialized heuristic exists. Extremal optimization (EO) falls into the latter category, adaptable to a wide range of combinatorial optimizations problems rather than crafted for a specific application. In close analogy to the Bak-Sneppen model of SOC, the EO algorithm proceeds as follows for the case of graph bi-partitioning: \begin{enumerate} \item Initially, partition the $N$ points at will into two equal subsets. \item Rank each point $i$ according to its fitness, $\lambda_i=g_i/(g_i+b_i)$, where $g_i$ is the number of (good) edges connecting $i$ to points within the same subset, and $b_i$ is the number of (bad) edges connecting $i$ to the other subset. If point $i$ has no connections at all ($g_i=b_i=0$), let $\lambda_i=1$. \item Pick the least fit point, {\em i.e.\/}, the point (from either subset) with the smallest $\lambda_i\in [0,1]$. Pick a second point at random from the other subset, and interchange these two points so that each one is in the opposite subset from where it started. \item Repeat at (2) for a preset number of times [assume $O(N)$ updates]. \end{enumerate} \noindent The result of an EO run is defined as the best (minimum cutsize) configuration seen so far. All that is necessary to keep track of, then, is the current configuration and the best so far. EO, like simulated annealing (SA) and genetic algorithms (GA), is inspired by observations of physical systems [for a comparison of SA and GA, see e. g. (de Groot 1991)]. However, SA emulates the behavior of frustrated systems in thermal equilibrium: if one couples such a system to a heat bath of adjustable temperature, by cooling the system slowly one may come close to attaining a state of minimal energy. SA accepts or rejects local changes to a configuration according to the Metropolis algorithm (Metropolis 1953) at a given temperature, enforcing equilibrium dynamics (``detailed balance'') and requiring a carefully tuned ``temperature schedule''. In contrast, EO takes the system far from equilibrium: it applies no decision criteria, and all new configurations are accepted indiscriminately. It may appear that EO's results would resemble an ineffective random search. But in fact, by persistent selection against the worst fitnesses, one quickly approaches near-optimal solutions. At the same time, significant fluctuations still remain at late run-times (unlike in SA), crossing sizable barriers to access new regions in configuration space, as shown in Fig.~\ref{runtime}. EO and genetic algorithms are equally contrasted. GAs keep track of entire ``gene pools'' of solutions from which to select and ``breed'' an improved generation of global approximations. By comparison, EO operates only with local updates on a single copy of the system, with improvements achieved instead by elimination of the bad. \begin{figure} \vskip 2.2truein \special{psfile=fig2.ps angle=-90 hoffset=-10 voffset=180 hscale=35 vscale=30} \caption{Evolution of the cutsize during an extremal optimization run on an $N=500$ geometric graph with $\alpha=5$ (see Fig.~\protect\ref{geograph}). The shaded area marks the range of cutsizes explored in the respective time bins. The best cutsize ever found is 2, which is visited repeatedly in this run. In contrast to simulated annealing, which has large fluctuations in early stages of the run and then converges much later, extremal optimization quickly approaches a stage where broadly distributed fluctuations allow it to probe many local optima. In this run, a random initial partition was used, and the runtime on a 200MHz Pentium was 9sec. } \label{runtime} \end{figure} Further improvements may be obtained through a slight modification of the EO procedure. Step (2) of the algorithm establishes a fitness rank for all points, going from rank $n=1$ for the worst fitness $\lambda$ to rank $n=N$ for the best. (For points with degenerate values of $\lambda$, the ranks may be assigned in random order.) Now relax step (3) so that the points to be interchanged are both chosen from a probability distribution over the rank order: from each subset, we pick a point having rank $n$ with probability $P(n)\propto n^{-\tau},~1\leq n\leq N$. The choice of a power-law distribution for $P(n)$ ensures that no regime of fitness gets excluded from further evolution, since $P(n)$ varies in a gradual, scale-free manner over rank. Universally, for a wide range of graphs, we obtain best results for $\tau\approx 1.2-1.6$. What is the physical meaning of an optimal value for $\tau$? If $\tau$ is too small, we often dislodge already well-adapted points of high rank: ``good'' results get destroyed too frequently and the progress of the search becomes undirected. On the other hand, if $\tau$ is too large, the process approaches a deterministic local search and gets stuck near a local optimum of poor quality. At the optimal value of $\tau$, the more fit components of the solution are allowed to survive, without the search being too narrow. Our numerical studies have indicated that the best choice for $\tau$ is closely related to a transition from ergodic to non-ergodic behavior, with optimal performance of EO obtained near the edge of ergodicity. To evaluate EO, we tested the algorithm on a testbed of well-studied large graphs\footnote{These instances are available at\newline http://userwww.service.emory.edu/\~{}sboettc/graphs.html} discussed in (Hendrickson 1996, Merz 1998). Table~\ref{tab1} summarizes EO's results on these, using 30 runs of at most $200N$ update steps (in several cases far fewer were necessary; see below). On the first four large graphs, SA's performance is extremely poor; we therefore substitute results given in (Hendrickson 1996) using a variety of specialized heuristics. EO significantly improves upon these cutsizes, though at longer runtimes. The best results to date on the graphs are due to various GAs (Merz 1998). EO reproduces all of these cutsizes, displaying an increasing runtime advantage as $N$ increases. On the final four graphs, for which no GA results were available, EO matches or dramatically improves upon SA's cutsizes. And although increasing $\alpha$ generally slows down EO and speeds up SA, EO's runtime is still nearly competitive with SA's on the high-connectivity {\em Nasa\/} graphs. Several factors account for EO's speed. First of all, in step (1) we employ a simple ``greedy'' start to form the initial partition, clustering connected points into the same partition from a random seed. This helps EO to succeed rapidly. By contrast, greedy initialization improves the performance of SA only for the smallest and sparsest graphs. Second of all, in step (2) we use a stochastic sorting process to accelerate the algorithm. At each update step, instead of perfectly ordering the fitnesses $\lambda_i$, we arrange them on an ordered binary tree called a ``heap''. We then select members from the heap such that {\em on average\/}, the actual rank selection approximates $P(n)\sim n^{-\tau}$. This stochastic rank sorting introduces a runtime factor of only $\alpha\log{N}$ per update step. Finally, EO requires significantly fewer update steps (Fig.~\ref{runtime}) than, say, a complete SA temperature schedule. The quality of our large $N$ results confirms that $O(N)$ update steps are indeed sufficient for convergence. In the case of the {\em Nasa\/} graphs, only $30N$ update steps (rather than the full $200N$) were in fact required for EO to reach its best results, and in the case of the {\em Brack2\/} graph, only $2N$ steps were required. \begin{table}[t] \caption{Best cutsizes and runtimes for our testbed of graphs. EO and SA results are from our runs (SA parameters as determined by Johnson {\em et al.\/} (Johnson 1989)), using a 200MHz Pentium. GA results are from Merz and Freisleben (Merz 1998), using a 300MHz Pentium. Comparison data for three of the large graphs are due to results from heuristics by Hendrickson (Hendrickson 1996), using a 50MHz Sparc20. } \begin{center} \begin{tabular}{lr@{\ }lr@{\ }lr@{\ }l} Graph & \multicolumn{2}{c}{EO} & \multicolumn{2}{c}{GA} & \multicolumn{2}{c}{heuristics} \\ \hline {\em Hammond\/} & 90 & (42s) & 90 & (1s) & 97 & (8s) \\ \multicolumn{7}{l}{\quad($N=4720$; $\alpha=5.8$)} \\ {\em Barth5\/} & 139 & (64s) & 139 & (44s) & 146 & (28s) \\ \multicolumn{7}{l}{\quad($N=15606$; $\alpha=5.8$)} \\ {\em Brack2\/} & 731 & (12s) & 731 & (255s) & \multicolumn{2}{c}{---} \\ \multicolumn{7}{l}{\quad($N=62632$; $\alpha=11.7$)} \\ {\em Ocean\/} & 464 & (200s) & 464 & (1200s) & 499 & (38s) \\ \multicolumn{7}{l}{\quad($N=143437$; $\alpha=5.7$)} \\ \hline \hline Graph &&& \multicolumn{2}{c}{EO} & \multicolumn{2}{c}{SA} \\ \hline {\em Nasa1824\/} &&& 739 & (6s) & 739 & (3s)\\ \multicolumn{7}{l}{\quad($N=1824$; $\alpha=20.5$)} \\ {\em Nasa2146\/} &&& 870 & (10s) & 870 & (2s)\\ \multicolumn{7}{l}{\quad($N=2146$; $\alpha=32.7$)} \\ {\em Nasa4704\/} &&& 1292 & (15s) & 1292 & (13s)\\ \multicolumn{7}{l}{\quad($N=4704$; $\alpha=21.3$)} \\ {\em Stufe10\/} &&& 51 & (180s) & 371 & (200s) \\ \multicolumn{7}{l}{\quad($N=24010$; $\alpha=3.8$)} \\ \end{tabular} \end{center} \label{tab1} \end{table} \section{Optimizing near Critical Points} Further comparison of EO and SA, averaged over a large sample of a particular type of graph, shows EO to be especially useful near critical points (Boettcher, to appear). It has been observed that many optimization problems exhibit critical points delimiting ``easy'' phases of a generally hard problem (Cheeseman 1991). Near such a critical point, finding solutions becomes particularly difficult for local search methods that explore some neighborhood in configuration space starting from an existing state. Near-optimal solutions become widely separated with diverging barrier heights between them. It is not surprising that equilibrium search methods based on heat-bath techniques like SA are not particularly successful here (Binder 1987). In contrast, the driven dynamics of EO does not possess any temperature control parameters that could increasingly limit the scale of fluctuations. A non-equilibrium approach like EO thus provides a general-purpose optimization method that is complementary to SA, which would be expected to freeze quickly into a poor local optimum ``where the {\em really\/} hard problems are'' (Cheeseman 1991). \begin{figure} \vskip 2.0truein \special{psfile=geomerror.ps angle=-90 hoffset=-10 voffset=180 hscale=35 vscale=30} \caption{Plot of SA's error relative to the best result found on geometric graphs, as a function of the mean connectivity $\alpha$. } \label{error} \end{figure} \begin{figure} \vskip 2.0truein \special{psfile=geomscal.ps angle=-90 hoffset=-10 voffset=180 hscale=35 vscale=30} \caption{Scaling plot of the data from EO according to Eq.~\protect\ref{scaleq} for geometric graphs, as a function of the mean connectivity $\alpha$. The scaling parameters and the fit are as discussed in the text. } \label{scaling} \end{figure} As an example, we explore this critical point for the equal partitioning of geometric graphs, as a function of their connectivity.\footnote{More results of this study, including many different types of graphs, can be found in (Boettcher, to appear).} It is hopeless to obtain reliable benchmarks for the exact optimal partition of large graphs. Instead, by averaging over many instances we can try to reproduce well-known results from the percolation properties of this class of graphs. For instance, when the average connectivity $\alpha$ of a geometric graph is much below $\alpha_{\rm crit}\approx4.5$, the percolation threshold found for these graphs (Balberg 1985), the graph most likely consists of many small clusters. These can easily be sorted into equal sized partitions with vanishing cutsize, at a cost of at most O($N^2$). When, on the other hand, the connectivity is large, the graph is dense and almost homogeneous with many near-optimal solutions in close proximity. But for connectivities near $\alpha_{\rm crit}$, a ``percolating'' cluster of size O($N$) appears with very widely separated minima (see Fig.~\ref{geograph}), making both the decision problem and the actual search very costly (Cheeseman 1991). We have generated geometric graphs of connectivities between $\alpha=4$ and $\alpha=10$ (by varying the threshold distance $d$ below which points are connected), at $N=500$, 1000, 2000, 4000, 8000, and 16000. For each $\alpha$ we generated 16 different instances of graphs, identical for SA and EO. We performed 32 optimization runs for each method on each instance. On each run, we used a different random seed to establish an initial partition of the points. SA was run using the algorithm developed by Johnson {\em et al.\/} (Johnson 1989) for this case, but with a temperature length four times longer, to improve results. EO was run for $200N$ update steps to produce a comparable runtime. For each method, we have taken only the best result from all runs on a given instance. We average those best results, for a particular connectivity $\alpha$, to obtain the mean cutsize for that method as a function of $\alpha$ and $N$. To compare EO and SA, we determine the relative error of SA with respect to the best result found by either method (most often by EO!) for $\alpha\geq\alpha_{\rm crit}$. Fig.~\ref{error} suggests that the error of SA diverges about linearly with increasing $N$, near $\alpha_{\rm crit}$. For the data obtained with EO, we make an Ansatz \begin{eqnarray} \langle m_{\rm opt}\rangle\sim N^\nu\left(\alpha-\alpha_{\rm 0}\right)^\beta \label{scaleq} \end{eqnarray} with $\nu=0.6$, in order to scale the data for all $N$ onto a single curve (see Fig.~\ref{scaling}). The remaining parameters are established according to a data fit, yielding $\alpha_0=4.1$ and $\beta=1.4$. The fact that $\alpha_0 < \alpha_{\rm crit}$ indicates that below the percolation threshold EO's cutsizes are already non-vanishing, and so even EO does not always find optimal partitions there. \section{Extremal Optimization of the TSP} In the graph partitioning problem, the implementation of EO is particularly straightforward. The concept of fitness, however, is equally meaningful in any optimization problem whose cost function can be decomposed into $N$ equivalent degrees of freedom. Thus, EO may be applied to many other NP-hard problems, even those where the choice of quantities for the fitness function, as well as the choice of elementary move, is less clear than in graph partitioning. One case where these choices are far from obvious is the traveling salesman problem. Even so, we have found there that EO presents a challenge to more finely tuned methods (Boettcher, submitted). In the traveling salesman problem (TSP), $N$ points (``cities'') are given, and every pair of cities $i$ and $j$ is separated by a distance $d_{ij}$. The problem is to connect the cities using the {\em shortest\/} closed ``tour'', passing through each city exactly once. For our purposes, take the $N\times N$ distance matrix $d_{ij}$ to be symmetric. Its entries could be the Euclidean distances between cities in a plane --- or alternatively, random numbers drawn from some distribution, making the problem non-Euclidean. (The former case might correspond to a business traveler trying to minimize driving time; the latter to a traveler trying to minimize expenses on a string of airline flights, whose prices certainly do not obey triangle inequalities!) For the TSP, we implement EO in the following way. Consider each city $i$ as a degree of freedom, with a fitness based on the two links emerging from it. Ideally, a city would want to be connected to its first and second nearest neighbor, but is often ``frustrated'' by the competition of other cities, causing it to be connected instead to (say) its $p$th and $q$th neighbors, $1\leq p\neq q\leq N-1$. Let us define the fitness of city $i$ to be $\lambda_i=3/(p_i+q_i)$, so that $\lambda_i=1$ in the ideal case. Defining a move class (step (3) in EO's algorithm) is more difficult for the TSP than for graph partitioning, since the constraint of a closed tour requires an update procedure that changes several links at once. One possibility, used by SA among other local search methods, is a ``two-change'' rearrangement of a pair of non-adjacent segments in an existing tour. There are $O(N^2)$ possible choices for a two-change. Most of these, however, lead to even worse results. For EO, it would not be sufficient to select two independent cities of poor fitness from the rank list, as the resulting two-change would destroy more good links than it creates. Instead, let us select one city $i$ according to its fitness rank $n_i$, using the distribution $P(n)\sim n^{-\tau}$ as before, and eliminate the longer of the two links emerging from it. Then, reconnect $i$ to a close neighbor, using the {\em same\/} distribution function $P(n)$ as for the rank list of fitnesses, but now applied instead to a rank list of $i$'s neighbors ($n=1$ for first neighbor, $n=2$ for second neighbor, and so on). Finally, to form a valid closed tour, one of the old links to the new (neighbor) city must be replaced; there is a unique way of doing so. For the optimal choice of $\tau$, this move class allows us the opportunity to produce many good neighborhood connections, while maintaining enough fluctuations to explore the configuration space. \begin{table}[t] \caption{Best tour-lengths found for the Euclidean (top) and the random-distance TSP (bottom). Results for each value of $N$ are averaged over 10 instances, using on each instance an exact algorithm (except for $N=256$ Euclidean where none was available), the best-of-ten EO runs, and the best-of-ten SA runs. Euclidean tour-lengths are rescaled by $1/\sqrt{N}$.} \begin{center} \begin{tabular}{rrrr} $N$ & Exact & EO$_{10}$ & SA$_{10}$\\ \hline Euclidean\qquad 16& 0.71453& 0.71453& 0.71453\\ 32& 0.72185& 0.72237& 0.72185\\ 64& 0.72476& 0.72749& 0.72648\\ 128& 0.72024& 0.72792& 0.72395\\ 256& \qquad --- & 0.72707& 0.71854\\ \hline Rand.\ Dist.\qquad 16& 1.9368& 1.9368& 1.9368\\ 32& 2.1941& 2.1989& 2.1953\\ 64& 2.0771& 2.0915& 2.1656\\ 128& 2.0097& 2.0728& 2.3451\\ 256& 2.0625& 2.1912& 2.7803 \end{tabular} \end{center} \label{tab2} \end{table} We performed simulations at $N=16$, 32, 64, 128, and 256, in each case generating ten random instances for both the Euclidean and non-Euclidean TSP. The Euclidean case consisted of $N$ points placed at random in the unit square with periodic boundary conditions; the non-Euclidean case consisted of a symmetric $N\times N$ distance matrix with elements drawn randomly from a uniform distribution on the unit interval. On each instance we ran both EO and SA, selecting for both methods the best of 10 runs from random initial conditions. EO used $\tau=4$ (Eucl.) and $\tau=4.4$ (non-Eucl.), with $16N^2$ update steps. SA used an annealing schedule with $\Delta T/T=0.9$ and temperature length $32N^2$. The results are given in Table~\ref{tab2}, along with baseline results using an exact algorithm. While the EO results trail those of SA by up to about 1\% in the Euclidean case, EO significantly outperforms SA for the non-Euclidean (random distance) TSP. Surprisingly, using increased run times (longer temperature schedules) diminishes rather than improves SA's performance in the latter case. Finally, note that one would not expect a general method such as EO to be competitive here with specialized optimization algorithms designed particularly with the TSP in mind. But remarkably, EO's performance in both the Euclidean and non-Euclidean cases --- within several percent of optimality for $N\le 256$ --- places it not far behind the leading specially-crafted TSP heuristics (Johnson 1997). \section{Extremal Optimization and Learning} Our results therefore indicate that a simple extremal optimization approach based on self-organizing dynamics can outperform state-of-the-art (and far more complicated or finely tuned) general-purpose algorithms on hard optimization problems. Based on its success on the generic and broadly applicable graph partitioning problem, as well as on the TSP, we believe the concept will be applicable to numerous other NP-hard problems. It is worth stressing that the rank ordering approach employed by EO is inherently non-equilibrium. Such an approach could not, for instance, be used to enhance SA, whose temperature schedule requires equilibrium conditions. This rank ordering serves as a sort of ``memory'', allowing EO to retain well-adapted pieces of a solution. In this respect it mirrors one of the crucial properties noted in the Bak-Sneppen model (Boettcher 1996). At the same time, EO maintains enough flexibility to explore further reaches of the configuration space and to ``change its mind''. Its success at this complex task provides motivation for the use of extremal dynamics to model mechanisms such as learning, as has been suggested recently to explain the high degree of adaptation observed in the brain (Chialvo 1999). \subsubsection*{References} C.~J.~Alpert and A.~B.~Kahng, {\em Integration: the VLSI Journal\/} {\bf 19}, 1 (1995). P.~Bak, {\em How Nature Works\/} (Springer, New York, 1996). P.~Bak, C.~Tang, and K.~Wiesenfeld, {\em Phys. Rev. Lett.\/} {\bf 59}, 381 (1987). P.~Bak and K.~Sneppen, {\em Phys. Rev. Lett.\/} {\bf 71}, 4083 (1993). I.~Balberg, Phys. Rev. B {\bf 31}, R4053 (1985). K.~Binder, {\em Applications of the Monte Carlo Method in Statistical Physics\/}, K.~Binder, Ed. (Springer, Berlin, 1987). S.~Boettcher and M.~Paczuski, {\em Phys. Rev. E\/} {\bf 54}, 1082 (1996), and {\em Phys. Rev. Lett.\/} {\bf 79}, 889 (1997). S.~Boettcher, {\em J.~Phys.~A:~Math.~Gen\/}, to appear; available at http://xxx.lanl.gov/abs/cond-mat/9901353. S.~Boettcher and A.~G.~Percus, submitted to \hfil\break {\em Artificial~Intelligence\/};~available~at \hfil\break http://xxx.lanl.gov/abs/cond-mat/9901351. V.~{\v C}erny, {\em J.~Optimization Theory Appl.\/} {\bf 45}, 41 (1985). P.~Cheeseman, B.~Kanefsky, and W.~M.~Taylor, in {\em Proc. of IJCAI-91}, J.~Mylopoulos and R.~Rediter, Eds. (Morgan Kaufmann, San Mateo, CA, 1991), pp.\ 331--337. D.~R.~Chialvo and P.~Bak, {\em J. Neurosci.\/}, to appear. C.~Darwin, {\em The Origin of Species by Means of Natural Selection\/} (Murray, London, 1859). A.~E.~Dunlop and B.~W.~Kernighan, {\em IEEE Trans. on Computer-Aided Design\/} {\bf CAD--4}, 92 (1985). M.~R.~Garey and D.~S.~Johnson, {\em Computers and Intractability: A Guide to the Theory of NP-Completeness\/} (Freeman, New York, 1979). S.~J.~Gould and N.~Eldridge, {\em Paleobiology\/} {\bf 3}, 115--151 (1977). C. de Groot, D. Wuertz, K. H. Hoffmann, {\em Lecture Notes in Computer Science\/} {\bf 496}, 445-454 (1991). B.~A.~Hendrickson and R.~Leland, in {\em Proceedings of the 1995 ACM/IEEE Supercomputing Conference (Supercomputing '95)\/}, San Diego, CA, December 3--8, 1995 (ACM Press, New York, 1996). J.~Holland, {\em Adaptation in Natural and Artificial Systems\/} (University of Michigan Press, Ann Arbor, 1975). D.~S.~Johnson, C.~R.~Aragon, L.~A.~McGeoch, and C.~Schevon, {\em Operations Research\/} {\bf 37}, 865 (1989). D.~S.~Johnson and L.~A.~McGeoch, in {\em Local Search in Combinatorial Optimization\/}, E.~H.~L.~Aarts and J.~K.~Lenstra, Eds. (Wiley, New York, 1997), chap.~8. S.~Kirkpatrick, C.~D.~Gelatt, and M.~P.~Vecchi, {\em Science\/} {\bf 220}, 671 (1983). B. B. Mandelbrot, {\em The Fractal Geometry of Nature\/} (Freeman, New York, 1983). P.~Merz and B.~Freisleben, in {\em Lecture Notes in Computer Science: Parallel Problem Solving From Nature --- PPSN V\/}, A.~E.~Eiben, T.~B\"ack, M.~Schoenauer, and H.-P.~Schwefel, Eds. (Springer, Berlin, 1998), vol. 1498, pp.\ 765--774, and {\em Technical Report No. TR--98--01\/} (Department of Electrical Engineering and Computer Science, University of Siegen, Siegen, Germany, 1998), available at http://www.informatik.uni-siegen.de/\~{}pmerz/publications.html. N.~Metropolis, A.~W.~Rosenbluth, M.~N.~Rosenbluth, A.~H.~Teller, and E.~Teller, {\em J.~Chem. Phys.\/} {\bf 21}, 1087 (1953). M.~Paczuski, S.~Maslov, and P.~Bak, {\em Phys. Rev. E\/} {\bf 53}, 414 (1996). I. Rodriguez-Iturbe and A. Rinaldo, {\em Fractal River Basins: Chance and Self-Organization\/} (Cambridge, New York, 1997). K.~Sneppen, P.~Bak, H.~Flyvbjerg, and M.~H.~Jensen, {\em Proc. Natl. Acad. Sci.\/} {\bf 92}, 5209 (1995). \end{document}
\section*{Acknowledgment} The author appreciatively acknowledges that this research is partially supported by Chinese Nature Science Foundation, by Chinese Ministry of Science \& Technology under the State Key Project of Basic Research on Rare Earth, and by Chinese Ministry of Education.
\section{Introduction} Metal cluster (MC) is a bound system consisting of atoms of some metal. The amount of atoms can vary from a few to many thousands. Some MC, mainly of alkali (Li, K, Na, ...) and noble (Ag, Au, ...) metals, demonstrate a striking similarity to atomic nuclei (see reviews$^{1-5}$). In these clusters the valence electrons are {\it weakly} coupled to the ions and, like nucleons in nuclei, are not strongly localized. The mean free path of valence electrons is of the same order of magnitude as the size of the cluster. This favors the valence electrons to form a mean field of the same kind as in nuclei (with the similar shell structure and magic numbers). In addition to the mean field, MC demonstrate other similarities with atomic nuclei: deformation in the case of open shells, variety of giant resonances (GR), fission, etc.. As a result, many theoretical ideas and methods of nuclear physics can, after a certain modification, be applied to MC \cite {Ne92,Br93,Ne_Dub}. This review is devoted to collective oscillations of valence electrons in MC. Valence electrons can be considered as the counterparts of nucleons in nuclei, and their oscillations as the counterparts of nuclear GR. Investigation of GR in MC is interesting in two aspects: it allows to understand deeper general properties of collective modes in finite Fermi systems and, simultaneously, allows to study striking peculiarities of MC. GR in clusters and atomic nuclei are well overlapped. However, some specific properties of MC cause considerable differences in the behavior of GR in these two systems. For example: the Coulomb interaction and the ''spill-out'' effect provide a specific dependence of GR properties on the mass number; the negligible character of the spin-orbital interaction leads to the decoupling of spin and orbital magnetic modes; clusters can have much more particles (atoms) than nuclei, which favors very strong orbital magnetic resonances; for most of the clusters the role of the ionic subsystem is important; at different temperatures MC can be in solid, liquid and even ''boiling'' phases, which greatly influences GR properties; characteristics of GR vary considerably whether the clusters are charged or neutral, free or embedded to a substrate, pure or with impurities atoms, etc.. Our consideration will be limited by certain physical conditions. -- i) The modern techniques allow to fabricate atomic clusters from atoms of about any element of the periodic table. However, the conception of the mean field for valence electrons is realized only for a minority, -- mainly for clusters of alkali and noble metals and, in a less extent, for neighboring elements. So, we should limit ourselves by this MC region. -- ii) In some alkali metals (Na and K) the ionic lattice can, to good accuracy, be replaced by a uniform distribution of the positive charge over cluster's volume. This is so-called jellium approximation which greatly simplifies the analysis and calculations. This approximation is enough for the description of many properties of alkali MC and will widely be used in the review. However, it often fails beyond Na and K and then a more explicit treatment of the ionic structure is necessary \cite{Br93,Bre_94rev}. -- iii) The ionic subsystem is supposed to be ''frozen''. In the review only collective oscillations of valence electrons will be considered. -- iv) The validity of jellium approximation is supported by temperature fluctuations of ions, which smooth ion positions. It fails in the low temperature region (approximately at $T<100$ K) where the explicit treatment of the ionic structure is important. At too high temperatures ($T>1000$ K) the quantum shells of the mean field are washed out, what establishes an upper limit for our considerations. We will consider GR in a temperature interval between these extreme cases. \section{Theoretical Grounds} \label{sec:the} Due to the similarity between MC and nuclei, many models of nuclear theory have been applied to study MC\cite{Ne92,Br93}. In due time, some of them have been introduced to nuclear physics from solid body field and then subsequently modified to describe {\it finite} Fermi systems. Now they turn out to be useful for clusters. In particular, a large variety of the RPA methods have been adopted , scaling from simple versions, like the sum rule approach$^{6-9}$ and the local RPA \cite{RB90,ReGB_AP96}, to sophisticated full RPA models, like time-dependent Hartree-Fock \cite{GJ92} and time-dependent local density approximation (TD-LDA)$^{13-26}$ (for a more complete list of citation see Refs.\cite{Ne92,Br93,Rubrev}). The simple models can describe the gross structure of GR but not the fragmentation of the collective strength. The full RPA models can describe the fragmentation but are very time consuming. The last shortcoming becomes crucial for deformed and large spherical clusters where the number of particles, and thus the size of the configuration space, is very large. In this connection, the intermediate class of the models, the RPA with {\it separable} residual forces (SRPA), seems to be very promising$^{7,27-35}$. The separable ansatz allows one to turn the RPA matrix into a simple dispersion relation. This drastically simplifies the eigenvalue problem preserving, at the same time, the main advantage of the full RPA to describe the fragmentation of the collective strength. The SRPA version derived in Refs.$^{27-33}$ provides the accuracy of full RPA calculations \cite {KNRS_EPJ98}, can be applied to systems of any shape \cite {Ne_ZPD,Ne_PRC,NK_Tsu,NKK_Pra}, and allows to treat GR in MC and atomic nuclei on the same microscopic footing \cite{NK_PS,Ne_PRC,NKK_Pra}. The results obtained within this SRPA version will be widely used in the review as illustrative examples. For the description of collective oscillations, the SRPA, and most of the other models, exploit, as a starting point, the Kohn-Sham energy functional \cite{KS65,GL76} for a system of $N_e$ valence electrons: \[ E\{n({\bf r},t),m({\bf r},t),\tau ({\bf r},t)\} =1/2\int \tau ({\bf r},t)d% {\bf r}+\int v_{xc}(n({\bf r},t),m({\bf r},t))d{\bf r} \] \begin{equation} +1/2\int \int \frac{(n({\bf r},t)-n_i({\bf r}))(n({\bf r_1},t)-n_i({\bf r_1}% ))}{|{\bf r}-{\bf r_1}|}d{\bf r}d{\bf r_1}, \end{equation} which includes the kinetic energy, the exchange-correlation term in the local density approximation (LDA) \cite{GL76,VWN80} and the Coulomb interaction, respectively. Here, $n({\bf r},t)=n({\bf r},t)\uparrow +n({\bf r% },t)\downarrow =\sum\nolimits_l|{\phi }_l({\bf r},t)|^2$, $m({\bf r},t)=n(% {\bf r},t)\uparrow -n({\bf r},t)\downarrow $ and $\tau ({\bf r}% ,t)=\sum\nolimits_l|\bigtriangledown {\phi }_l({\bf r},t)|^2$ are the density, magnetization density (z-component) and kinetic energy density of valence electrons, respectively; $n_i({\bf r})$ is the ionic density in the jellium approximation; $\phi _l({\bf r},t)$ is a single-particle wave function. The convention $e=m_e=\hbar =1$ is used. The functional (1) can have additional terms if the ionic structure is treated beyond the jellium approximation. The time-dependent single-particle Hamiltonian is obtained as \begin{equation} H({\bf r},t)\phi _l({\bf r},t)=\frac{\delta E}{\delta \phi _l^{*}({\bf r},t)}% . \end{equation} In the small-amplitude limit of a collective motion, the densities can be written as $n({\bf r},t)=n_{{0}}({\bf r})+\delta n({\bf r},t)$ and $m({\bf r}% ,t)=m_{{0}}({\bf r})+\delta m({\bf r},t)$ where $n_0({\bf r})$ and $m_0({\bf % r})$ are the static ground state densities ($m_0({\bf r})=0$ in spherical and unpolarized clusters) and the values $\delta n({\bf r},t)$ and $\delta m(% {\bf r},t)$ are small time-dependent density variations (transition densities). Then, in the linear approximation to the density variations, the Hamiltonian (2) is a sum of the static and dynamical parts. The static part \begin{equation} H_0({\bf r})=T+V_0({\bf r})=-\frac{\triangle }2+(\frac{dv_{xc}}{dn})_{n=n_{{0% }},m=m_{{0}}}+\int \frac{n_{{0}}({\bf r_1})-n_i({\bf r_1})}{|{\bf r}-{\bf r_1% }|}d{\bf r_1} \end{equation} constitutes the Kohn-Sham single-particle potential. It can be approximated with a good accuracy by phenomenological potentials, such as the harmonic oscillator\cite{LiS89_ZPD} (for small spherical MC), Nillson-Clemenger\cite {C85,RFB95} (for deformed MC) or Woods-Saxon \cite{Ne_PRA,NHM90,FP96} (for spherical and deformed MC . In the electric channel, the dynamic part of the Hamiltonian (residual interaction) has the form \begin{equation} \delta H({\bf r},t)=(\frac{d^2v_{xc}}{dn^2})_{n=n_0}\delta n({\bf r},t)+\int \frac{\delta n({\bf r_1},t)}{|{\bf r}-{\bf r_1}|}d{\bf r_1}. \label{eq:ELres} \end{equation} The dominant term here is the Coulomb interaction. The residual interaction in this channel is always positive (repulsive) and shifts the unperturbed electrical multipole strength from the typical particle-hole (ph) values $% \omega _{ph}=0.9-1.5$ eV to higher energies 2.6-3.2 eV. In the spin channel, the dynamical part is \begin{equation} \delta H({\bf r},t)=(\frac{d^2v_{xc}}{dndm})_{n=n_0,m=m_0}\delta n({\bf r}% ,t)\delta m({\bf r},t). \label{eq:MLres} \end{equation} Here, the residual interaction is determined by the exchange-correlation term. It is always negative (attractive) and shifts the unperturbed magnetic multipole strength from $\omega _{ph}=0.9-1.5$ eV to lower energies 0.2-0.8 eV. The residual interactions (4) and (5) are typical for the TD-LDA calculation scheme. In what follows we will mainly consider clusters constituted from monovalent atoms, like alkali metals, for which the numbers of valence electrons and atoms coincide, $N_e=N$. \section{Electric Dipole Giant Resonance (E1 GR)} Unlike nuclei, where different kinds of GR are well investigated both experimentally and theoretically, our knowledge in clusters is mainly limited by the electric dipole resonance (dipole plasmon). Experimentally the E1 GR has been observed in a variety of clusters: small and large, spherical and deformed, neutral and charged, hot and cooled (see, for example, Refs.$^{43-49}$). As a rule, the photoabsorption cross section was measured by methods of the depletion spectroscopy. For other GR ($EL(L\ne 1)$, $ML$) there are only theoretical predictions$^{6-9,28,30,32,50,51}$. Physical interpretations of E1 GR in clusters and nuclei are very similar: while in nuclei it is caused by translations of neutrons against and protons, then in clusters it is a result of translations of the valence electron against ions \cite{YanB_PR91}. In spite of this similarity, the dipole resonance in clusters exhibits many interesting peculiarities which will be discussed below. \subsection{Energy of E1 GR: Step by Step} The description of E1 energy for clusters is a rather complicated task. For example, while in nuclei its energy depends on the mass number as $A^{-1/3}$% , in clusters the E1-energy can both decrease (Ag clusters) and increase (alkali MC) with the number of atoms. Let us consider this important characteristic step by step. {\bf Step one: Mie frequency and spill-out effect}. In the simplest approximation, MC can be considered as a {\it classical} metallic drop. Then, the E1-energy is described by Mie expression \cite{Mie}: $\omega _{Mie}=\omega _p/\sqrt{3}$ where $\omega _p$ is the plasma frequency. For Na clusters $\omega _{Mie}=3.41eV$. This value is much higher than the experimental E1-energy which is 2.5-2.8 eV for spherical Na clusters with $% N<100$. The agreement with the experiment is considerably improved if we take into account the {\it quantum} spill-out effect. This effect means that since the valence electrons are quantum entities, they are not well localized and so, unlike the classical ionic jellium, can be partly {\it spilled out } beyond the jellium boundary. In principle, this effect takes place in any two-component quantum system including atomic nuclei and atoms (a ''neutron skin'' in small nuclei is a relevant example). With the spill-out, the E1 energy in MC is described as \cite{LiS89_ZPD} \begin{equation} \omega _{E1}=\omega _{Mie}(1-\frac 12\frac{\delta N_e}{N_e}) \end{equation} where $\delta N_e$ is the number of spilled out valence electrons. As a result, the discrepancy with the experiment reduces to 0.2-0.3 eV. The spill-out effect allows to explain the increase of the E1-energy with N, observed in alkali MC. The value $\delta N_e$ decreases with the size (for example, $\delta N_e=1.5(19\%)$ and $9.5(7\%)$in $Na_8$ and $Na_{138}$, respectively \cite{Ar_NC89}) leading to the corresponding increase in the E1-energy. {\bf Step two: beyond jellium approximation, ionic structure, local and nonlocal effects.} The remaining discrepancy can be removed in a large extent by the explicit treatment of the ionic subsystem. First of all, we should take into account that ions are not the points but have a size. Inside this size, ion core electrons (ICE) (do not confuse them with the valence electrons) screen the pure Coulomb interaction of ions and valence electrons. To take into account this screening, the atomic pseudopotentials (PP) are used (see, for instance, \cite{BG95,CCG95,YB96,BHS82}). They allow to describe correctly the spectrum of valence electrons in isolated atoms {\it without} the solution of the complicated many-body atomic task. Being a sum of contributions of ICE with different orbital momenta, PP have {\it % local} (s-electrons) and {\it nonlocal} (p and d electrons) parts \cite {BHS82}. To avoid dealing with nonlocal functions, the Pseudo-Hamiltonians (PH) were introduced as the next simplifying step \cite{BCC89,L96}. PH, being derived from PP, lead to less involved (but with the same accuracy) calculations since, unlike the PP, they treat the nonlocality only through the differential operators. Folding atomic PH with jellium, one gets PH for {\it atomic clusters}$^{19-21}$. PH have the additional advantage to be easily incorporated to the common calculation schemes. As compared to the conventional Kohn-Sham Hamiltonian, PH include the additional local and two non-local (the orbital contribution and the effective mass) terms. As is seen from Figure~1, in K clusters (the same for Na) the nonlocal contributions are negligible and the local term is enough to get good description of the E1-energy\cite {KNRS_EPJ98}. This is not the case for Li clusters, where only both, local and nonlocal, contributions provide the agreement with the experiment \cite {KNRS_EPJ98}. In some studies (see, e.g., Ref.~\cite{BG95}) the ICE effects are taken into account together with some averaged treatment of the ionic arrays in a cluster. The latter leads to an additional, but rather moderate, redshift of the E1-energy. {\bf Step three: direct dynamical ICE contribution.} The ICE effects discussed above are realized through the change of the single-particle characteristics with the subsequent renormalization of the residual interaction. Besides this {\it indirect} way, the ICE can {\it directly} influence the dynamics and thus lead to new peculiarities of the E1 GR. This can be well demonstrated for Ag clusters where, like atomic nuclei and {\it unlike} alkali MC, the E1-energy {\it % decreases} with a size \cite{Ti_92GR}. The physics behind is that in these clusters the energy of ICE excitations is comparable to the E1-energy. The coupling of these two modes additionally screens the interaction of valence electrons with ions (direct dynamical ICE contribution) and finally causes the redshift (decrease) of the E1-energy \cite{SR97}. This effect is mainly of a volume character and so intensifies with a cluster size. As a result, the E1 energy in Ag clusters decreases with N. This tendency overpowers the opposite one caused by the spill-out effect. In alkali clusters, where the ICE excitations have energies much higher than the E1 GR, the direct dynamical ICE contribution can be neglected and the evolution of the E1-energy with the cluster size is determined mainly by the spill-out effect. \section{Landau Damping and Width of E1 GR} The main physical mechanisms forming the plasmon width are the thermal fluctuations of a cluster shape and the Landau damping (RPA fragmentation of the collective strength) \cite {Br93,YVB93,KNRS_EPJ98,R_SRPA,YB_APNY91,BL95,MonRei_temp}. The relative contributions of these two mechanisms change with a cluster size. As is seen from Figure~2, in small clusters, like $Na_{21}^{+}$, where the Landau damping is negligible, the thermal fluctuations determine about all the width. In clusters of moderate size, like $Na_{59}^{+}$, the Landau damping is stronger and greatly contributes to the width. This is especially the case for deformed clusters. In large clusters, like $Na_{441}^{+}$, the Landau damping is weaker though its contribution to the width remains to be considerable. Figure~2 (bottom) shows that the Landau damping is closely related with the shell structure \cite{KNRS_EPJ98}. In $Na_{21}^{+}$ the dipole plasmon lies in the wide gap between the bunches of $\Delta {\cal N}$% =1 and $\Delta {\cal N}$=3 particle-hole (ph) states and remains almost unperturbed as a collective peak. With increasing the cluster size, the resonance approaches the bunch $\Delta {\cal N}$=3 and, in $Na_{59}^{+}$, is already interferes with ph states of this bunch, which leads to the considerable Landau damping. For larger clusters, the plasmon runs to the swamp of ph states. This leads to a general trend of increasing the width which is, however, overlaid by sizeable fluctuations \cite{KNRS_EPJ98,R_SRPA}% . But here a further mechanism comes into play: the coupling between the resonance and ph states fades away due to increasing mismatch of $\Delta {\cal N}$=1 ph configurations (which mainly generate the plasmon) and surrounding ph states with much larger values of $\Delta {\cal N}$. This finally leads to a decrease of the plasmon width $\propto N_e^{-1/3}$ estimated analytically in the wall formula \cite{YB_APNY91} and tested in the RPA calculations \cite{R_SRPA}. The Landau damping in MC with $N<40$ is rather sensitive to cluster charge: being strongest in negatively charged ones (anions), the Landau damping is considerably reduced while passing to neutral and then to positively charged clusters (cations) \cite{Yan_CPL}. This effect is caused by a strong dependence of the single-particle potential depth $V_0$ on the cluster charge. In anions the potential is shallow ($V_0\simeq -2$ eV), the energy gaps between $\Delta {\cal N}$ bunches are very smooth and, so, there are good conditions for a sizeable Landau damping (see discussion above). In neutral clusters and more in cations, the potential depth is increased to about -7 eV, the gaps between $\Delta {\cal N}$ bunches in the ph spectrum become more distinctive, which weakens the Landau damping. \section{Temperature Effects} In most of experiments with GR in MC, the typical cluster temperature is estimated to be in the interval 300-900 K which corresponds to the thermal energy $kT=0.03-0.09$ eV. At these temperatures, ions behave as classical particles and quantum properties of the cluster are mainly determined by valence electrons. This can be easily proved \cite{B90} by using the uncertainty relation $\Delta x\Delta p\geq \hbar $. This relation gives lower bounds for the momentum and energy of a particle in a system: $\Delta p=\hbar /\Delta x$ and $\Delta E=(\Delta p)^2/2m$, respectively. Taking $% \Delta x\leq 1.5\AA $ (the diameter of $Na_{20}$) for both ions and valence electrons, one gets \[ \Delta E_e\geq 0.16eV\qquad \mbox{for electrons,} \] \[ \Delta E_i\geq 10^{-4}eV\quad \mbox{for ions.} \] The energy of a quantum motion of valence electrons considerably exceeds the thermal energy, which favors their quantum behavior. The opposite situation takes place for ions which, therefore, should exhibit the classical behavior. Such result takes place because the ionic mass is much larger than the electron one. The difference in ionic and electron masses leads to other interesting consequence. Namely, almost all the thermal energy is contained in the ionic subsystem. Valence electrons are embedded to the thermal ionic bath. So, unlike atomic nuclei, MC represent the case of the canonical ensemble. The bulk melting points for K, Na and Li are $T_b=336$, 371 and 452 K, respectively. This means that most of the measurements for GR in MC have been done for clusters in a liquid-like phase. As was mentioned above, in small clusters, thermal shape fluctuations provide the dominate contribution to the plasmon width. It is interesting that, while in nuclei these fluctuations are mainly of a quadrupole form, in MC they are mainly octupole \cite{MonRei_temp}. The reason is that spherical and neighboring MC are soft to the octupole deformation. Photoexcitation is a rapid process in the ionic time scale. So, every response of a cluster represents its instantaneous shape and the experimental cross section gives a properly weighted response of all allowed shapes \cite{BL95}. The higher the temperature, the larger the plasmon width and the smaller the plasmon energy. The temperature shift is estimated as about $1\%$ of the plasmon energy per 100 K \cite{Bre91,HB67}. It can be explained by the effectively increase of the cluster size with temperature. The larger the size, the bigger the static dipole polarizability which is expressed through the cluster radius as $\alpha _{E1}=R^3$. The polarizability is connected with the plasmon energy through the inverse sum rule, $\alpha _{E1}=2m_{-1}\simeq <E1>^2/\omega _{E1}$. So, the higher the temperature, the larger $\alpha _{E1}$ and, consequently, the smaller $\omega _{E1}$. Recent experiments show that at sufficiently low temperatures the gross-structure of the E1 GR drastically changes \cite{Na10_PRL}. For example, the axially deformed cluster $Na_{11}^{+}$ at 380 K demonstrates the typical two-peak spectrum determined by the deformation splitting of E1 GR. At 35 K the same resonance exhibits much more complicated structure including at least 6 well-distinguished peaks. This structure reflects the ionic arrangement which is not washed out at so low temperature by variations of ions. In this case, the jellium approximation is not valid and models based on this cannot be applied. The E1 GR in small clusters at low temperature seems to be best described by {\it ab initio} quantum-chemical calculations \cite{BK_Pra}. \section{E1 GR in Deformed Clusters} Like nuclei, MC with open shells have quadrupole deformation$^{45-49,64-68}$% . There are experimental indications of both prolate and oblate axial quadrupole shapes, as well as of $\gamma $-deformation$^{45-49}$. In the framework of different methods (Strutinski's shell correction method, ultimate jellium model, etc.) hexadecapole and octupole deformations as well as high isomerism have been predicted$^{64-68}$. Rather strong quadrupole, hexadecapole and octupole deformations should take place at least up to MC with $N\sim 700$ \cite{FP96Er}. Like in nuclei, E1 GR in axially deformed MC exhibits the deformation splitting in two peaks (see Figure~3). The right peak is about twice larger than the left one in prolate clusters (see $Na_{11}^{+}$, $% Na_{15}^{+}$, $Na_{27}^{+}$) and, vice versa, in oblate clusters (see $% Na_{35}^{+}$). Most of MC are deformed. But getting an experimental information on a cluster shape, even in the simplest case of a quadrupole deformation, is rather nontrivial problem. In nuclei rotational bands serve as source of such information. In principle, deformed clusters can rotate. But, due to a large value of the moment of inertia, rotational energies are very small and, being of the same order of magnitude as the thermal energy, fail to be observed. In this connection, the splitting of E1 GR in deformed clusters is now a {\it single direct} manifestation of quadrupole deformation and the main source of the information about it. \section{Multipole GR, Asymptotic Trends, Restoring Forces} So far, the depletion spectroscopy methods (photoabsorption and photofragmentation) were mainly exploited for observation of E1 GR in MC \cite{He93}. The other reactions $((e,e^{\prime }),(\gamma ,\gamma ^{\prime })$ and etc.) are not yet sufficiently developed, which impedes the observation of other GR. The similar situation took place in nuclear physics in early seventies. For this reason an investigation of EL GR with $L\ne 1$ is yet limited to theoretical predictions$^{6-9,28,30,32,50,51}$. In Figure 4, E2 and E3 GR in spherical $Na_{59}^{+}$, calculated within the SRPA, are presented as typical examples. It is instructive to consider the main trends of electrical multipole giant resonance with the size (N) and multipolarity (L), and also the origin of the GR restoring forces. Such analysis has been done within the sum rule approach (SRA) in Ref.\cite{Se_SRA_PRB89}. In the jellium approximation for valence electrons, $n_{{0}}(r)=n_i(r)=n^{+}\theta (r-R)$ (the spill-out effect is neglected), the energy of $EL(L\ne 0)$ GR can be written as \cite {Se_SRA_PRB89} \begin{equation} \omega _{EL}=\sqrt{\frac{m_3}{m_1}}=\hbar \sqrt{\frac 23(2L+1)(L-1)\frac{% \beta _F^2}{R^2}+\omega _p^2\frac L{2L+1}} \label{eq:omEL} \end{equation} where $m_1=\sum_iB(EL,gr\to i)\omega _i$ and $m_3=\sum_iB(EL,gr\to i)\omega _i^3$ are the sum rules, $\beta _F=(3/5)^{1/2}(3\pi ^2)^{1/3}\frac 1mn_0^{1/3}$ and R is the radius of a cluster. The first term in Eq. \ref {eq:omEL} is the contribution of the kinetic energy (the similar expression have been obtained earlier in Ref. \cite{NS80}). The second term is determined by the Coulomb interaction between valence electrons (ee) and valence electrons and ions (ei). Eq. \ref{eq:omEL} shows that E1 GR is determined only by the Coulomb interaction. In the limit of large R, one has \begin{equation} \omega _{EL}\to \hbar \omega _p\sqrt{\frac L{2L+1}}. \end{equation} The larger $L$, the higher the excitation energy of the GR. In general, due to the first term in Eq. \ref{eq:omEL}, the energy of $EL(L\ne 0,1)$ GR is decreased with N. For low L in small clusters this tendency is changed by the spill-out effect. The separate analysis for E0 GR gives the increase of the E0 energy with N to the limit $\omega _{E0}\to \hbar \omega _p$. \begin{table}[tbp] \caption{Relative contributions to $m_3$ for $Na_{92}$: kinetic energy $% (m_3(T))$, exchange and correlations $(m_3(xc))$, electron-electron interaction $(m_3(ee)$, electron-ion interaction $(m_3(ei))$ and total Coulomb interaction $(m_3(C)=m_3(ee)+m_3(ei))$ \protect\cite{Se_SRA_PRB89}. } \label{tab:m3EL}\vspace{0.4cm} \par \begin{center} {\footnotesize \begin{tabular}{|c|c|c|c|c|c|} \hline L & $m_3(T)$ & $m_3(xc)$ & $m_3(ee)$ & $m_3(ei)$ & $m_3(C)$ \\ \hline 1 & 0 & 0 & 0 & 1 & 1 \\ 2 & 0.08 & 0 & -0.77 & 1.69 & 0.92 \\ 5 & 0.51 & 0 & -2.29 & 2.78 & 0.49 \\ \hline \end{tabular} } \end{center} \end{table} It is seen from Eq. \ref{eq:omEL} that the value $m_3$ has the meaning of a restoring force \cite{RB90}. In Table \ref{tab:m3EL} the contributions to $% m_3$ from different terms of the Kohn-Sham functional (1) are presented. It is seen that the restoring force for E1 GR is determined by the electron-ion interaction (ei) only. With increasing L, the electron-electron contribution (ee) raises and starts to compensate the (ei) Coulomb part. Simultaneously, the kinetic energy term grows. For high L, the contribution of the total Coulomb interaction goes to zero and all the restoring force is determined by the kinetic energy. The purely volume exchange correlation term (xc) which within the LDA depends only on the electron density does not contribute to $m_3$. The restoring force should not be confused with the residual interaction. As is seen from Eq. \ref{eq:ELres}, the residual interaction, unlike the restoring force, has only the (ee)- and (xc)-terms for any L (where the (ee)-term dominates). \section{Anharmonicity and Multiphonon GR} How much harmonic are the GR in metal clusters? How strong is the mixing of one and two phonon states? Investigations performed within different approaches$^{71-73}$ give contradictory answers for one-phonon GR. While the shell-model calculations found for E1 GR in $Na_{20}$ some signals of anharmonicity \cite{KML_94}, other studies predict the harmonic character for M2(spin-dipole) and EL GR \cite{Cat_93,CRS97}. It should be noted that these studies have been done for rather small clusters with $N\le 20$. In this size region the GR energy lies safely below the lowest 2p-2h configurations, what does not favor the anharmonic effects. This picture can change in larger clusters where GR approach the region of 2p-2h configurations. The calculations \cite{Cat_93} predict a noticeable anharmonicity for most of {\it double} (two-phonon) GR placed at 8-15 eV. These GR exhibit a weak mixing with one-phonon states but a considerable fragmentation between two-phonon configurations. Most strong effect is expected for some $0^{+}$ double GR, for example, for $(1^{-}\otimes 1^{-})_{0^{+}}$ in $Na_{21}^{+}$. With the appearance of new experimental techniques allowing investigation of multiple GR these predictions are quite important. The techniques use non-intense femtosecond lasers \cite{S98} or exploit collisions of a cluster with highly charged ions \cite{G97}. Quite recently the multiple GR constructed from 3-4 dipole plasmons has been observed in $Na_{93}^{+}$ \cite {S98}. \section{Magnetic GR} Like in atoms, the spin-orbital interaction in metal clusters is negligible and thus spin and orbital collective magnetic modes are well decoupled. The separation of these two modes in MC is easier than in nuclei. \subsection{Spin-Multipole GR} Magnetic multipole resonances (ML) of spin character caused by the external field $Q_L=\sum_{j=1}^Nr_j^LY_{L0}\sigma _j^z$ were studied within the SRA and RPA \cite {LiC_96spin,SL_97spin,SBBN_93spin,MCSRe_96spin}. For $L=1$ the operator $Q_1\sim \sum_{j=1}^Nz_j\sigma _j^z$ provides the opposite shifts of spin-up and spin-down electrons in z-direction. Unlike the electric resonances , the residual interaction for ML GR is defined only by the exchange and correlations ((xc)-term) since only the (xc)-term depends on the magnetization density (see Eq. \ref{eq:MLres}) . Therefore the study of ML resonances can provide a valuable information about (xc)-effects in clusters. Approximating the electron density by the expression $n_0=n_{00}/(1+exp((r-R)/a)$, one gets the energy for ML GR \cite{SL_97spin} \begin{eqnarray} \omega _{ML} &=&\sqrt{\frac{m_3}{m_1}}=\hbar [\frac 25(2L+1)(L-1)\frac{\beta _F^2}{R^2}+\frac{e^2}{m}4\pi aL\frac{n_0}{R^{2L-1}} \nonumber \\ &+&\frac 1mL(v_{xc}^{(02)}(n_0,m_0)-v_{xc}^{(20)}(n_0,m_0))\frac{n_0}{6aR}% ]^{1/2} \label{eq:omML} \end{eqnarray} where $v_{xc}^{(pq)}(n_0,m_0)=\frac{d^p}{dn^p}\frac{d^q}{dm^q}% v_{xc}^{(pq)}(n,m)\mid _{(n=n_0,m=m_0)}$. For other notation see Eq. \ref {eq:omEL}. As compared to Eq. \ref{eq:omEL} for EL GR, Eq. (9), was derived taking into account the spill-out effect. Furthermore, due to the presence of the spin in the operator $Q_L$, the (xc)-term now contributes to $m_3$, unlike the case of EL GR. However, the exchange contributions (Pauli principle) to $v_{xc}^{20}$ and $v_{xc}^{02}$ are the same and then, only correlation effects enter Eq. \ref{eq:omML}. The energies of spin ML resonances decrease with N and go to zero for large sizes. The larger $L$, the higher the GR energy. The behavior of ML GR much depends on the diffuseness parameter $a$. \begin{table}[t] \caption{ Relative contributions to $m_3$ for $Na_{92}$ and $Na_{912}$ : kinetic energy $(m_3(T))$, correlation $(m_3(c))$ and total Coulomb interaction $(m_3(C))$. The data are extracted from the Fig. 1 of Ref. \protect\cite{SL_97spin}. } \label{tab:m3ML}\vspace{0.4cm} \par \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|} \hline & \multicolumn{3}{|c|}{$Na_{92}$} & \multicolumn{3}{|c|}{$Na_{912}$} \\ \cline{2-7} L & $m_3(T)$ & $m_3(c)$ & $m_3(C)$ & $m_3(T)$ & $m_3(c)$ & $m_3(C)$ \\ \hline 1 & 0 & 0.77 & 0.23 & 0 & 0.30 & 0.70 \\ 2 & 0.50 & 0.49 & 0.01 & 0.31 & 0.68 & 0.01 \\ 5 & 0.76 & 0.20 & 0.04 & 0.59 & 0.38 & 0.03 \\ \hline \end{tabular} \end{center} \end{table} Table \ref{tab:m3ML} demonstrates that the restoring force for spin-multipole GR differs from the electrical GR case. Namely, the contribution of correlations\cite{VWN80} dominates for $L=1$ and 2 and remains to be considerable for larger $L$. The correlation term includes long-range RPA correlations$^{76-78}$, short-range correlations \cite{LST94} and others. The correlations greatly influence both static and dynamical characteristics of MC$^{37,38,76-78}$ and their investigation is very important. \subsection{Orbital GR} Since the number of atoms in MC can be much more than the number of nucleons in nuclei, much larger values of the single-particle orbital moment can be achieved what can give origin to very strong orbital magnetic multipole resonances, orbital ML GR. That is, clusters, as nuclei, can exhibit orbital ML GR like,``scissors'', twist mode, etc., nevertheless with a much stronger strength \cite{LiS89_ZPD,Ba94}. Investigations of the specific low-energy orbital M1 GR, which can exist only in {\it deformed} clusters demonstrated that it can serve as a good indicator of the cluster quadrupole deformation $^{7,81-83}$. Indeed, in some cases the deformation splitting of E1 GR is washed out by other effects and is not enough distinctive to get a reliable information on cluster deformation. Then the orbital M1 GR can be used for this aim. Macroscopically, this resonance is treated as small-angle rigid rotations of the ellipsoid of valence electrons against the ionic ellipsoid. Such collective mode was shown to be coupled with the quadrupole component, $% \bigtriangledown (yz)$, of the displacement field \cite{LiS89_ZPD,LS89}. The orbital M1 GR has the counterpart in deformed nuclei, well known as the ''scissors'' mode\cite{IP78}. The latter describes the rotations of the neutron ellipsoid against the proton one. The orbital M1 GR is represented by $K^\pi =1^{+}$ states ($K$ is the angular-momentum projection) with a low excitation energy and strong M1 transitions to the ground state. For Na clusters these characteristics are estimated as \cite{LiS89_ZPD,LS89} $\omega _{M1}=4.6\beta _2N_e^{-1/3}(1+5\frac{\omega _0}{\omega _p})^{-1/2}$ eV and $B(M1)=1.1\beta _2N_e^{4/3}\mu _b^2$ where $% \beta _2$ is the deformation parameter, $B(M1)$ is the reduced transition probability and $\omega _0$ is the harmonic oscillator frequency. Both $% \omega _{M1}$ and $B(M1)$ are proportional to the deformation parameter and so the orbital M1 GR survives only in deformed clusters. \begin{table}[t] \caption{The excitation energy and strength (within the interval 0-1 eV) of orbital M1 GR, calculated within the SRPA \protect\cite{NKS_Nas,NKSI_SN}. See the text for notation. } \label{tab:OM1}\vspace{0.4cm} \par \begin{center} \begin{tabular}{|c|c|c|c|c|c|} \hline & $Na^+_{15}$ & $Na^+_{27}$ & $Na^+_{35}$ & $Na^+_{119}$ & $Na^+_{295}$ \\ \hline $\beta_2$ & 0.32 & 0.23 & -0.23 & 0.25 & 0.24 \\ $\omega _{M1}$, eV & 0.63 & 0.29 & 0.35 & 0.26 & 0.21 \\ $B(M1), \mu_b^2$ & 27 & 56 & 41 & 229 & 757 \\ \hline \end{tabular} \end{center} \end{table} The results of the first realistic RPA calculations for orbital M1 GR \cite {NKS_Nas,NKSI_SN} are given in Table 3. It is seen that this resonance has low excitation energies. The most remarkable result is that already in clusters with about 300 atoms, the orbital M1 GR strength reaches very high values, 700-800 $\mu _b^2$. This GR is described in detail in Ref.\cite {NKSI_SN} of the present Proceedings. \section{Other GR in Atomic Clusters} As compared to nuclei, atomic clusters provide many specific manifestations of E1 GR. For example, clusters embedded in a dielectric matrix demonstrate a strong screening effect: the matrix screens the residual interaction between valence electrons in a cluster, which results in the considerable decrease of E1-energy \cite{RS93}. In mixed and coated clusters the impurity (or coated) atoms much influence both the ground state and properties of E1 GR (see, e.g. Refs.$^{86-88}$). In the fullerene $C_{60}$, two E1 GR are known as determined by weakly bonded $\pi $ electrons and strongly bonded $\sigma $ electrons (see, e.g. Ref. \cite{GL93_C60}). $^3He$ and $^4He$ clusters representing collections of fermions ($^3He$ atoms) and bosons ($^4He$ atoms), respectively, should be mentioned. In $% ^3He $ clusters just $^3He$ atoms (not valence electrons) form a mean field with quantum shells \cite{B90,S91,WR93}. These clusters are characterized by strong surface effects. Unlike nuclei and MC, $^3He$ clusters represent the case of {\it one-component} Fermi-system and, so, have no E1 GR. At the same time, the study of other EL GR reveals new possibilities, for instance, the comparison of the GR properties in Fermi ($^3He$ clusters) \cite{WR92,S91_He} and Bose ($^4He$ clusters)\cite{CS90} systems. \section*{Summary} Giant resonances in atomic clusters have been observed. Being much similar to their counterparts in atomic nuclei, GR in MC demonstrate, at the same time, numerous exciting peculiarities. The unique situation takes place now in many-body physics where, in addition to atoms and atomic nuclei, a new family of small Fermi systems (MC, fullerenes, $He^3$ clusters, quantum dots) appears. This greatly enlarges our possibilities in many-body studies. All mentioned systems possess, in a different extent, a mean field with quantum shells. It should be noted that atomic clusters are attractive both for fundamental studies and practical applications \cite{Ne_JINR}. Last achievements (creation of new materials, machinning superhard surfaces, creation of extremely large energy densities in a matter, catalysis, microelectronics, microcomputering, etc.) show that, due to atomic clusters, one may expect in a recent future a remarkable progress in many high-tech fields. \section{Acknowledgments} We are grateful to M. Schmidt and H. Haberland for communication the experimental results and to the Organizing Committee of the Workshop for the financial support of the attendance. The work was also partly supported by CAPES (V.O.N.) and FINEP Brasil (V.O.N. and F.F.S.C.). \section*{References}
\section*{I. Introduction} Recently, a method based on successive canonical transformations has been used to obtain exact solution of the Schr\"odinger equation \begin{equation} i\frac{d}{dt}|\psi(t)\rangle=H(t)|\psi(t)\rangle \label{sch-eq} \end{equation} for a class of dipole Hamiltonians \cite{pla97-1,pra97-1,jmp97-2} and time-dependent harmonic oscillators \cite{pra97-2}. For these systems the Hamiltonian is a nondegenrate Hermitian operator. The purpose of the present article is to extend the application of this method to the cases where the Hamiltonian is non-Hermitian and involves degenerate eigenvalues. Non-Hermitian Hamiltonians have been used to model a variety of physical systems involving decaying states, \cite{nhh-application}. The solution of the Schr\"odinger equation for a time-dependent two-level non-Hermitian Hamiltonian has been considered in Refs.~\cite{da-to-mi,kv-pu}. Another motivations for the study of the Schr\"odinger equation for a time-dependent non-Hermitian Hamiltonian is the fact that the solution of every linear ordinary differential equation (ODE) may be reduced to the solution of a system of first order linear ODEs which can be written in the form of the time-dependent Schr\"odinger equation~(\ref{sch-eq}) or alternatively \begin{eqnarray} |\psi(t)\rangle&=&U(t)|\psi(0)\rangle\;, \label{psi=}\\ i\frac{d}{dt}U(t)&=&H(t)U(t)\;, \label{sch-eq-u}\\ U(t)&=&1\;, \label{ini-condi} \end{eqnarray} where $U(t)$ is the evolution operator. For a general linear ODE the corresponding Hamiltonian $H(t)$ may be a non-Hermitian matrix with degenerate eigenvalues. The method of {\em adiabatic product expansion} developed in Refs.~\cite{pla97-1,pra97-1} does not directly apply to quantum systems with non-Hermitian Hamiltonians. In this article we shall present a generalization of this method which applies to arbitrary (possibly) non-Hermitian Hamiltonians with degenerate as well as nondegenerate eigenvalues. The organization of the article is as follows. In section~II we review the basic results concerning the adiabatic approximation for degenerate and non-Hermitian Hamiltonians. In section~III we discuss the generalization of the method of adiabatic product expansion to these Hamiltonians. In section~IV we use the results of section~III to study the solution of the Schr\"odinger equation for a general nondegenerate non-Hermitian two-level Hamiltonian. In section~V, we apply the general results obtained in section~IV to treat the classical equation of motion for a harmonic oscillator with a time-dependent frequency. In section~VI, we discuss the application of the adiabatic product expansion to study the quadrupole interaction of a spin 1 particle with a time-dependent electric field $\vec{\cal E} =({\cal E}_1(t), {\cal E}_2(t),0)$. We show that the corresponding Hamiltonian which has a degenerate and a nondegenerate eigenvalue is canonically equivalent to a Hamiltonian which has only nondegenerate eigenvalues. Furthermore, we show that if the direction of the electric field depends in a particular way on its magnitude, then our method yields the exact solution of the Schr\"odinger equation. Finally we present our conclusions in section~VII. \section*{II. Adiabatic Approximation for Non-Hermitian Hamiltonians} Let $H=H[R]$ be a parametric Hamiltonian which depends on a set of real parameters $R=(R^1,R^2,\cdots,R^d)$ labelling the points of a smooth manifold $M$. Let $E_n[R]$ denote the eigenvalues of $H[R]$ and ${\cal H}_{n}[R]$ be the degeneracy subspace associated with $E_n[R]$. Let $\mbox{\footnotesize${\cal N}$}$ denote the degree of degeneracy of $E_n[R]$, i.e., the complex dimension of ${\cal H}_{n}[R]$. We shall assume that the spectrum of $H[R]$ is discrete and $\mbox{\footnotesize${\cal N}$}$ does not depend on $R$. Now let $|\psi_n,a;R\rangle$ and $|\phi_n,a;R\rangle$ form a complete biorthonormal basis of the Hilbert space \cite{wong,ga-wr}. This means that $|\psi_n,a;R\rangle$ with $a\in\{1,2,\cdots,\mbox{\footnotesize${\cal N}$}\}$ form a basis of ${\cal H}_{n}[R]$, in particular \begin{equation} H[R]|\psi_n,a;R\rangle=E_n[R]|\psi_n,a;R\rangle\;, \label{eg-va-eq-H} \end{equation} and $|\phi_n,a;R\rangle$ satisfy \begin{eqnarray} &&H[R]^\dagger|\phi_n,a;R\rangle=E_n^*[R]|\phi_n,a;R\rangle\;, \label{eg-va-eq-H*}\\ &&\langle\phi_m,b;R|\psi_n,a;R\rangle=\delta_{mn}\delta_{ab}\;, \label{biorthonormal}\\ &&\sum_{n}\sum_{a=1}^{\mbox{\footnotesize${\cal N}$}}|\psi_n,a;R\rangle\langle\phi_n,a;R|=1\;. \label{complete} \end{eqnarray} Next suppose that the parameters $R^i$ depend on time $t$, then $R(t)$ defines a curve ${\cal C}$ in the parameter space $M$, and the Hamiltonian, its eigenvalues and eigenvectors become time-dependent. In this case we use the notation $H(t):=H[R(t)]$, $E_n(t):=E_n[R(t)]$, $|\psi_n,a;t\rangle:= |\psi_n,a;R(t)\rangle$, and $|\phi_n,a;t\rangle:=|\phi_n,a;R(t)\rangle$. We shall assume that $E_n(t),~|\psi_n,a;t\rangle$ and $|\phi_n,a;t\rangle$ are smooth functions of $t$ and that during the evolution of the system the eigenvalues of the Hamiltonian do not cross, i.e., if $E_m(0)<E_n(0)$, then for all $t\in [0,\tau]$, $E_m(t)<E_n(t)$, where $\tau$ denotes the duration of the evolution of the system. Differentiating both sides of Eq.~(\ref{eg-va-eq-H}) with respect to $t$, taking the inner product of both sides of the resulting equation with $|\phi_m,b;t\rangle$, for arbitrary $m$ and $b$, and using Eqs.~(\ref{eg-va-eq-H}) -- (\ref{biorthonormal}), we have \begin{equation} [E_m(t)-E_n(t)]\langle\phi_m,b;t|\frac{d}{dt}|\psi_n,a;t\rangle+ \langle\phi_m,b;t|\dot H(t)|\psi_n,a;t\rangle -\delta_{mn}\delta_{ab}\dot E(t)=0\;. \label{1} \end{equation} Here a dot denotes differentiation with respect to $t$. For $m\neq n$, Eq.~(\ref{1}) reads \begin{equation} \langle\phi_m,b;t|\frac{d}{dt}|\psi_n,a;t\rangle=\frac{ \langle\phi_m,b;t|\dot H(t)|\psi_n,a;t\rangle}{E_n(t)-E_m(t)}~~~ {\rm for}~~~m\neq n. \label{2} \end{equation} Now let us express the solution of the Schr\"odinger equation~(\ref{sch-eq}) in the basis $\{|\psi_n,a;t\rangle\}$. Then \begin{equation} |\psi(t)\rangle=\sum_{n}\sum_{a=1}^{\mbox{\footnotesize${\cal N}$}}C_a^n(t)|\psi_n,a;t\rangle\;, \label{3} \end{equation} where $C_a^n(t)$ are complex coefficients. Substituting Eq.~(\ref{3}) in the Schr\"odinger equation~(\ref{sch-eq}), taking the inner product of both sides of the resulting equation with $|\phi_m,b;t\rangle$, and making use of Eqs.~(\ref{eg-va-eq-H}), (\ref{eg-va-eq-H*}), (\ref{biorthonormal}), and (\ref{2}), we find \begin{equation} i\dot C_b^m-E_m C_b^m+\sum_{a=1}^{\mbox{\footnotesize${\cal N}$}} i\langle\phi_m,b;t|\frac{d}{dt} \psi_m,a;t\rangle C_a^m=-i\sum_{n\neq m}\sum_{a=1}^{\mbox{\footnotesize${\cal N}$}} \frac{\langle\phi_m,b;t|\dot H(t)|\psi_n,a;t\rangle}{E_n(t)-E_m(t)}\;. \label{4} \end{equation} The special case of this equation with $\mbox{\footnotesize${\cal N}$}=1$, i.e., the nondegenerate case, has been originally derived by Garrison and Wright \cite{ga-wr} in their investigation of the adiabatic geometric phase \cite{berry1984} for non-Hermitian Hamiltonians \cite{ga-wr,da-mi-to,mi-si-ba-be,mo-he,p28}. If the right-hand side of Eq.~(\ref{4}) is negligible, then one says that the system undergoes an {\rm adiabatic evolution}, \cite{bo-fo,kato,pra97-1,ga-wr,ne-ra}. In this case, the equations for $C_a^n$ decouple and their solution is given by \begin{equation} C^n_a(t)=\sum_{b=1}^{\mbox{\footnotesize${\cal N}$}} K^n_{ab}(t)C^n_b(0)\;, \label{5} \end{equation} where $K_{ab}^n(t)$ are entries of the invertible matrix \begin{equation} K^n(t):=e^{-i\int_0^tE_n(s)ds}\;{\cal P} \exp \left[i\int_{R(0)}^{R(t)} {\cal A}^n[R]\right]\;, \label{6} \end{equation} ${\cal P}$ denotes the path-ordering operator, ${\cal A}^n$ is the matrix of one-forms with entries \begin{equation} {\cal A}^n_{ab}[R]:=i\langle\phi_n,a;R|d|\psi_n,b;R\rangle\;, \label{a=} \end{equation} $d$ stands for the exterior derivative with respect to $R^i$, and the line intergral in Eq.~(\ref{6}) is evaluated along the curve ${\cal C}$ defined by $R(t)$. If ${\cal C}$ is a closed curve in $M$, the Hamiltonian has a periodic time-dependence and the path-ordered exponential in Eq.~(\ref{6}), which takes the form \begin{equation} {\cal P}\exp \left[i\oint_{\cal C}{\cal A}^n[R]\right]\;, \label{gp} \end{equation} is the {\em non-Hermitian} analogue of the {\em non-Abelian adiabatic geometric phase} \cite{wi-ze}. Note that if the initial vector $|\psi(0)\rangle$ is an eigenvector of the initial Hamiltonian $H(0)$, then the adiabaticity of the evolution implies that $|\psi(t)\rangle$ is an eigenvector of $H(t)$ for all $t\in[0,\tau]$. In terms of the time-evolution operator $U(t)$ of Eq.~(\ref{sch-eq-u}) this is expressed by \begin{eqnarray} U(t)&\approx& U^{(0)}(t)\;,~~~{\rm where} \label{7}\\ U^{(0)}(t)&:=&\sum_{n}\sum_{a,b=1}^{\mbox{\footnotesize${\cal N}$}} K^n_{ab}(t) |\psi_n,a;t\rangle\langle\phi_n,b;0|\;. \label{u0} \end{eqnarray} One can easily show that $U^{(0)}(t)$ is invertible, and its inverse is given by \begin{equation} U^{(0)^{-1}}(t)=\sum_n\sum_{a,b=1}^{\mbox{\footnotesize${\cal N}$}} K^{n^{-1}}_{ab}(t) |\psi_n,a;0\rangle\langle\phi_n,b;t|\;, \label{u0-1} \end{equation} where $K^{n^{-1}}(t)$ is the inverse of $K^n(t)$. \section*{III. Adiabatic Canonical Transformations and the Generalized Adiabatic Product Expansion} Let $g(t)$ be an invertible linear operator acting on the Hilbert space. Then the transformations: \begin{eqnarray} |\psi(t)\rangle&\to& |\psi'(t)\rangle:=g(t)|\psi(t)\rangle\;, \label{psi-trans}\\ H(t)&\to& H'(t):=g(t)H(t)g(t)^{-1}-ig(t)\frac{d}{dt}g(t)^{-1}\;, \label{H-trans}\\ U(t)&\to&U'(t):=g(t)U(t)g(0)^{-1}\;, \label{u-trans} \end{eqnarray} leave the form of the Schr\"odinger equation invariant. We shall call such a transformation a {\em canonical transformation.} Now let us investigate the consequences of the canonical transformation defined by $g(t)=U^{(0)}(t)^{-1}$. We shall call this transformation the {\em adiabatic canonical transformation.} Denoting the transformed Hamiltonian $H'$ by $H^{(1)}$, we have \begin{eqnarray} H^{(1)}(t)&=&\sum_{n,m\neq n}\sum_{a=1}^{\mbox{\footnotesize${\cal N}$}} \sum_{b=1}^{\cal M}H^{(1)^{nm}}_{ab}(t)|\psi_n,a;0\rangle\langle \phi_m,b;0|\;,~~~{\rm where} \label{H1}\\ H^{(1){nm}}_{ab}(t)&:=&-K_{ac}^{n}(t)^{-1}A_{cd}^{nm}(t) K_{db}^m(t)\;,~~~{\rm and}~~~ A_{cd}^{nm}(t):=i\langle\phi_n,c;t|\frac{d}{dt}|\psi_m,d;t\rangle\;. \label{H1-ab} \end{eqnarray} Because $g(0)=U^{(0)}(0)^{-1}=1$, the transformed evolution operator is given by \begin{equation} U'(t)=U^{(0)}(t)^{-1} U(t) \;. \label{u'=} \end{equation} Clearly if the adiabatic approximation is valid, $H^{(1)}(t)\approx 0$ and $U'(t)\approx 1$. Let us suppose that $H^{(1)}(t)$ has a discrete spectrum and denote by $E^{(1)}_{n_1}(t)$ and $\mbox{\footnotesize${\cal N}$}_1$ the eigenvalues of $H^{(1)}(t)$ and their degree of degeneracy. Furthermore, let $\{|\psi_{n_1}^{(1)},a_1;t\rangle, |\phi_n^{(1)},a_1;t\rangle\}$ be a biorthonormal eigenbasis of the Hilbert space, i.e., \begin{eqnarray} &&H^{(1)}(t)|\psi^{(1)}_{n_1},a_1;t\rangle=E^{(1)}_{n_1}[R] |\psi^{(1)}_{n_1},a_1;t\rangle\;,\nonumber\\ &&H^{(1)}(t)^\dagger|\phi^{(1)}_{n_1},a_1;t\rangle=E^{(1)*}_{n_1}(t) |\phi^{(1)}_{n_1},a_1;t\rangle\;,\nonumber\\ &&\langle\phi^{(1)}_{m_1},b_1;t|\psi^{(1)}_{n_1},a_1;t\rangle= \delta_{m_1n_1}\delta_{a_1b_1}\;,\nonumber\\ &&\sum_{n_1}\sum_{a_1=1}^{\mbox{\footnotesize${\cal N}$}_1}|\psi^{(1)}_{n_1},a_1;t\rangle \langle\phi^{(1)}_{n_1},a_1;t|=1\;.\nonumber \end{eqnarray} Then $H^{(1)}(t)$ shares the properties of the original Hamiltonian $H(t)$, and we can repeat the above analysis using $H^{(1)}(t)$ in place of $H(t)$. In this way the adiabatic approximation yields the approximate evolution operator \begin{equation} U^{(1)}(t)=\sum_{n_1}\sum_{a_1,b_1=1}^{\mbox{\footnotesize${\cal N}$}_1} K^{(1)n_1}_{a_1b_1}(t)|\psi_{n_1}^{(1)},a_1;t\rangle \langle\phi_{n_1}^{(1)},b_1;0| \label{u1=} \end{equation} for $H^{(1)}(t)$, where $K^{(1)n_1}_{a_1b_1}(t)$ are the entries of the matrix $K^{(1)n_1}$ obtained by replacing $E_n(t),~ |\psi_n,a;t\rangle$ and $|\phi_n,a;t\rangle$ in Eqs.~(\ref{6}) and (\ref{a=}) by $E_{n_1}^{(1)}(t)$, $|\psi_{n_1}^{(1)},a_1;t\rangle$, and $|\phi_{n_1}^{(1)},a_1;t\rangle$, respectively. Next we perform the adiabatic canonical transformation defined by $g(t)= U^{(1)}(t)^{-1}$. This leads to a transformed Hamiltonian $H^{(2)}(t)$ which is related to $H^{(1)}(t)$ according to Eqs.~(\ref{H1}) and (\ref{H1-ab}) with $K^n,~|\psi_n,a;t\rangle$ and $|\phi_n,b;t\rangle$ replaced by $K^{(1)^{n_1}},~|\psi_{n_1}^{(1)},a_1;t\rangle$ and $|\phi_{n_1}^{(1)},a_1;t\rangle$. The transformed evolution operator is given by \[U^{(1)}(t)^{-1}U^{(0)}(t)^{-1}U(t)\;.\] Repeating this procedure we obtain, after $N$ successive adiabatic canonical transformations, a transformed Hamiltonian $H^{(N)}(t)$ and a transformed evolution operator which is given by \[U^{(N-1)}(t)^{-1}U^{(N-2)}(t)^{-1}\cdots U^{(0)}(t)^{-1}U(t)\;.\] Here $U^{(\ell)}(t)$, with $\ell\in\{1,2,\cdots,N-1\}$, denotes the approximate evolution operator obtained by performing adiabatic approximation on the Hamiltonian $H^{(\ell)}(t)$. If for some $N$ the adiabatic approximation yields the exact solution of the Schr\"odinger equation for the Hamiltonian $H^{(N)}(t)$, then by construction $H^{(N+1)}(t)=0$ and $U^{(N+1)}(t)=1$. In this case, the original evolution operator is given by \begin{equation} U(t)=U^{(0)}(t)U^{(1)}(t)\cdots U^{(N)}(t)\;. \label{u=approx} \end{equation} If the adiabatic approximation fails for all $H^{(N)}(t)$, then there are two possibilities: \begin{itemize} \item[i)] one obtains an infinite product expansion for the evolution operator \begin{equation} U(t)=\prod_{\ell=0}^{\infty} U^{(\ell)}(t):= U^{(0)}(t)U^{(1)}(t)\cdots U^{(\ell)}(t)\cdots\;. \label{u=exact} \end{equation} In this case, one may view Eq.~(\ref{u=approx}) as a {\em generalization} of the adiabatic approximation. \item[ii)] one obtains $H^{(i)}(t)=H^{(j)}(t)$ for some $i$ and $j$ with $i\neq j$. In this case a direct application of the method of adiabatic product expansion does not produce a solution. However, as we shall see in the following section, sometimes it is possible to modify this method by combining the adiabatic canonical transformation with other canonical transformations, so that one obtains a finite or an infinite product expansion with distinct terms. \end{itemize} \section*{IV. Application to Two-Level Hamiltonians} Two-level nondegenerate Hamiltonians provide the simplest nontrivial quantum systems. This has been one of the main reasons for the study of these Hamiltonians since the early days of quantum mechanics. In this section we shall consider the most general nondegenerate two-level Hamiltonian which may or may not be Hermitian. In an arbitrary basis of the Hilbert space ($\relax\hbox{\kern.25em$\inbar\kern-.3em{\rm C}$}^2$), the Hamiltonian is given by a two-by-two complex matrix $\bar H$. One can perform a quantum canonical transformation (\ref{H-trans}) defined by $g(t)=\exp\{i\int_0^t [{\rm tr}~\bar H(s)] ds/2\}$ to map the Hamiltonian $\bar H$ to a traceless Hamiltonian of the form \begin{equation} H:=\left(\begin{array}{cc}a & b \\ c&-a \end{array}\right)\;, \label{2-level-H-traceless} \end{equation} where ${\rm tr}~ \bar H$ denotes the trace of $\bar H$, and $a=a(t),~b=b(t),~c=c(t)$ are complex-valued smooth functions of $t$. We can easily solve the eigenvalue problem for the Hamiltonian (\ref{2-level-H-traceless}). The eigenvalues are given by \begin{equation} E_1(t):=-E(t),~~~E_2(t):=E(t)\;,~~~{\rm where}~~~E:=\sqrt{a^2+bc}. \label{2-level-eg-va} \end{equation} We shall demand that during the time interval $[0,\tau]$ of interest $E\neq 0$, so that the eigenvalues are nondegenerate. In particular, no level crossings occur. Then a possible choice for a biorthonormal eigenbasis is \begin{eqnarray} &&|\psi_1;R\rangle=\left(\begin{array}{c} -b\\ a+E\end{array}\right)\,,~~~ |\psi_2;R\rangle=\left(\begin{array}{c} a+E\\ c\end{array}\right)\, \label{eg-vec}\\ &&|\phi_1;R\rangle=\frac{1}{N^*}\left(\begin{array}{c} -c^*\\ a^*+E^*\end{array}\right)\,,~~~ |\phi_2;R\rangle=\frac{1}{N^*}\left(\begin{array}{c} a^*+E^*\\ b*\end{array}\right)\;, \label{dual-eg-vec} \end{eqnarray} where $R=(a,b,c)$ and $N:=2E(a+E)$. Next we compute $U^{(0)}$ and $H^{(1)}$. Using Eqs.~(\ref{u0}), (\ref{6}), (\ref{a=}), and ({\ref{H1}), we find \begin{eqnarray} U^{(0)}(t)&=&K^1(t)|\psi_1;t\rangle\langle\phi_1;0|+ K^2(t)|\psi_2;t\rangle\langle\phi_2;0|\;, \label{u0-2}\\ H^{(1)}(t)&=&\xi(t)|\psi_1;0\rangle\langle\phi_2;0|+\zeta(t) |\psi_2;0\rangle\langle\phi_1;0|\;, \label{H1-2} \end{eqnarray} where \begin{eqnarray} K^1(t)&:=&K^1_{11}(t)=\exp\left( \frac{i\eta(t)}{2}- \int_{R(0)}^{R(t)}\left[(2E)^{-1}(da+dE+\frac{cdb}{a+E})\right] \right)\;,\nonumber\\ K^2(t)&:=&K^2_{11}(t)=\exp\left( \frac{-i\eta(t)}{2}- \int_{R(0)}^{R(t)}\left[(2E)^{-1}(da+dE+\frac{bdc}{a+E})\right] \right)\;,\nonumber\\ \eta(t)&:=&2\int_0^t E(s)ds\;, \label{eta}\\ \xi(t)&:=&H^{12}_{11}(t)=(-\frac{ie^{-2i\alpha(t)}}{2}) \left[1+\frac{a(t)}{E(t)}\right] \frac{d}{dt}\left[\frac{c(t)}{a(t)+E(t)}\right]\;, \label{xi}\\ \zeta(t)&:=&H^{21}_{11}(t)=(\frac{ie^{2i\alpha(t)}}{2}) \left[1+\frac{a(t)}{E(t)}\right] \frac{d}{dt}\left[\frac{b(t)}{a(t)+E(t)}\right]\;, \label{zeta}\\ \alpha(t)&:=&\frac{\eta(t)}{2}+\frac{i}{4}\int_{R(0)}^{R(t)}\frac{ cdb-bdc}{E(E+a)}\;. \label{alpha} \end{eqnarray} The transformed Hamiltonian has the following matrix expression \begin{equation} H^{(1)}(t)=\left(\begin{array}{cc} a^{(1)}(t) & b^{(1)}(t)\\ c^{(1)}(t) & -a^{(1)}(t)\end{array}\right)\;, \label{H1-2-matrix} \end{equation} where \begin{eqnarray} a^{(1)}(t)&:=&-\frac{b_0\xi(t)+c_0\zeta(t)}{2E_0}\;, \label{a1}\\ b^{(1)}(t)&:=&-\frac{b_0^2\xi(t)-(a_0+E_0)^2\zeta(t)}{2E_0(E_0+a_0)}\;, \label{b1}\\ c^{(1)}(t)&:=&-\frac{-(a_0+E_0)^2\xi(t)+c_0^2\zeta(t)}{2E_0(E_0+a_0)}\;, \label{c1}\\ a_0&:=&a(0)\;,~~~b_0:=b(0)\;,~~~c_0:=c(0)\;,~~~E_0:=E(0)\;,\nonumber \end{eqnarray} and we have used Eqs.~(\ref{eg-vec}), (\ref{dual-eg-vec}), and (\ref{H1-2}). Note that the transformed Hamiltonian $H^{(1)}(t)$ is traceless, and one can obtain $H^{(2)}(t)$ by substituting $a^{(1)}$ for $a$, $b^{(1)}$ for $b$, $c^{(1)}$ for $c$, and $E^{(1)}:=\sqrt{(a^{(1)})^2+ b^{(1)}c^{(1)}}$ for $E$ in Eqs.~(\ref{H1-2}), and (\ref{xi}) -- (\ref{alpha}). Clearly this can be repeated indefinitely, and one can compute $H^{(\ell)}$ for arbitrary $\ell$. The adiabatic approximation corresponds to the cases where the matrix elements of $H^{(1)}(t)$ can be neglected. As seen from Eqs.~(\ref{H1-2-matrix}) -- (\ref{c1}) this happens whenever both $\xi$ and $\zeta$ are negligible. One can also check that if only one of these quantities is negligible, then $H^{(1)}(t)$ is equal to the other times a constant matrix. This means that $H^{(1)}(t)$ has essentially stationary eigenvectors and the adiabatic approximation would yield the solution of the Schr\"odinger equation for $H^{(1)}$. In fact, it is not difficult to check that for the cases that either $\xi$ or $\zeta$ is negligible, $H^{(2)}(t)\approx 0$. In particular, setting $\xi=0$ or $\zeta=0$ implies $H^{(2)}(t)= 0$ and the evolution operator is given by \begin{equation} U(t)=U^{(0)}(t)U^{(1)}(t)\;. \label{u=uu} \end{equation} Therefore, the conditions $\xi=0$ and $\zeta=0$ each define a class of exactly solvable two-level systems. In view of Eqs.~(\ref{xi}) and (\ref{zeta}), these are \begin{itemize} \item[] Class 1: The two level systems for which $\frac{c}{a+E}=\mu =$ constant, or alternatively $c=\mu(\mu b+\sqrt{4a^2+\mu^2 b^2})/2$; \item[] Class 2: The two level systems for which $\frac{b}{a+E}=\nu=$ constant, or alternatively $c=b/\nu^2-a^2/b$. \end{itemize} In general $\xi$ and $\zeta$ do not vanish and the adiabatic product expansion does not terminate. There is also a special class of two-level systems for which the product expansion has a periodic structure in the sense of case (ii) of the preceding section. This is \begin{itemize} \item[] Class 3: The two level systems for which $a=0$. \end{itemize} Setting $a=0$ in Eqs.~(\ref{alpha}),~(\ref{eta}), (\ref{xi}), and (\ref{zeta}) and defining $f(t):=i\sqrt{c(t)/b(t)}$, we have \begin{eqnarray} \alpha(t)&=&\frac{\eta(t)}{2}+\frac{i}{4}\ln\left(\frac{c_0b(t)}{ b_0c(t)}\right)\;,~~~~ \eta(t)=2\int_0^t\sqrt{b(s)c(s)}ds\nonumber\\ \xi(t)&=& -\frac{f_0\dot f(t) e^{-i\eta(t)}}{2f(t)}\;,~~~~ \zeta(t)=\frac{\dot f(t) e^{i\eta(t)}}{2f_0f(t)}\;,\nonumber \end{eqnarray} where $f_0:=f(0)$. Substituting these equations in Eq.~(\ref{H1-2-matrix}), we obtain \begin{equation} H^{(1)}(t)=E^{(1)}(t) \left(\begin{array}{cc} \cos\eta(t)&f_0^{-1}\sin\eta(t)\\ f_0\sin\eta(t)&-\cos\eta(t)\end{array}\right)\;,~~{\rm where}~~ E^{(1)}(t)=\frac{i\dot f(t)}{2f(t)}\;. \label{H1-a=0} \end{equation} This Hamiltonian has two interesting properties. \begin{itemize} \item[1.] If $b_0=c_0$, then $f_0=i$ and \begin{equation} H^{(1)}(t) =E^{(1)}(t) \left[ \sin\eta(t)\sigma_2+ \cos\eta(t)\sigma_3\right]=E^{(1)}(t) e^{i\eta(t)\sigma_1/2} \sigma_3e^{-i\eta(t)\sigma_1/2}\;, \label{pauli} \end{equation} where $\sigma_i$ are Pauli matrices, and we have used the identity \begin{equation} e^{-i\varphi\sigma_i}\sigma_je^{i\varphi\sigma_i}= \cos(2\varphi)\sigma_j+\sin(2\varphi)\sum_{k=1}^3\epsilon_{ijk} \sigma_k\;,~~~~{\rm for}~~~i\neq k\;. \label{identity} \end{equation} In Eq.~(\ref{identity}), $\varphi$ is an arbitrary complex variable and $\epsilon_{ijk}$ is the totally anti-symmetric Levi Civita symbol with $\epsilon_{123}=1$. For the time periods during which $\sqrt{b(t)c(t)}$ is real, $\eta(t)$ is real, and the Hamiltonian (\ref{pauli}) is anti-Hermitian. In particular, its eigenvectors are orthogonal. Up to a factor of $i$ this Hamiltonian describes the interaction of a spin 1/2 magnetic dipole with a changing magnetic field. This system has a $SU(2)$ dynamical group \cite{su2,su,pla97-1,jmp97-2}. For the time periods during which $\sqrt{b(t)c(t)}$ is imaginary, $\eta(t)$ is imaginary, and up to a factor of $i$ the Hamiltonian (\ref{pauli}) describes a quantum system with a $SU(1,1)$ dynamical group. A Hermitian analogue of such a system is the time-dependent generalized harmonic oscillator \cite{su1-1,su,jackiw}. \item[2.] Performing the adiabatic canonical transformation on (\ref{pauli}), we arrive at the unexpected result \begin{equation} H^{(2)}(t)=H(t). \label{H2=H} \end{equation} Therefore, direct application of the method of adiabatic product expansion does not lead to a solution. \end{itemize} Next we shall describe a modification of the method of adiabatic product expansion which yields an infinite product expansion for the evolution operator of the Class~3 systems which involve distinct terms. Consider the transformed Hamiltonian (\ref{H1-a=0}). We can express this Hamiltonian using Eq.~(\ref{H1-2-matrix}) with \begin{equation} a^{(1)}(t)=E^{(1)}(t)\cos\eta(t)\;,~~~ b^{(1)}(t)=f_0^{-1}E^{(1)}(t)\sin\eta(t)\;,~~~ c^{(1)}(t)=f_0E^{(1)}(t)\sin\eta(t)\;. \label{a1-b1-c1} \end{equation} Although this Hamiltonian does not belong to Class~3, it can be canonically transformed to a Hamiltonian which belongs to Class~3, i.e., its diagonal matrix elements vanish. This transformation is defined by $g(t)=\exp\{i\int_0^ta^{(1)}(s)ds\sigma_3\} $. The corresponding transformed Hamiltonian is given by \begin{equation} H_1(t)=\left(\begin{array}{cc} 0 & b_1(t) \\ c_1(t) & 0 \end{array}\right) \;, \label{H_1} \end{equation} where \begin{equation} b_1(t):=b^{(1)}(t)e^{i\gamma_1(t)}\;,~~~ c_1(t):=c^{(1)}(t)e^{-i\gamma_1(t)}\;,~~ {\rm and}~~\gamma_1 (t):=2\int_0^ta^{(1)}(s)ds\;. \label{a_1-b_1-gamma_1} \end{equation} The evolution operator $U_1$ of $H_1$ is related to the evolution operator of the original Hamiltonian $H$ according to \begin{equation} U_1(t)=e^{i\int_0^ta^{(1)}(s)ds\sigma_3}~U^{(0)}(t)^\dagger ~U(t)\;, \label{u_1} \end{equation} where we have used Eq.~(\ref{u-trans}). Now since $H_1$ has the same form as $H$, we can repeat the above analysis using $H_1$ in place of $ H$. Performing an adiabatic canonical transformation on $H_1$ we obtain the transformed Hamiltonian \begin{equation} H_1^{(1)}(t)=\left(\begin{array}{cc} a_1^{(1)}(t) & b_1^{(1)}(t)\\ c_1^{(1)}(t) & -a_1^{(1)}(t)\end{array}\right)\;, \label{H1-2-matrix-2} \end{equation} where \begin{eqnarray} a_1^{(1)}(t)&:=&E_1^{(1)}(t)\cos\eta_1(t)\;,~~~ b_1^{(1)}(t):=f_{1,0}^{-1}E_1^{(1)}(t)\sin\eta_1(t)\;,\nonumber\\ c_1^{(1)}(t)&:=&f_{1,0}E_1^{(1)}(t)\sin\eta_1(t)\;,~~~ E_1^{(1)}(t):=E^{(1)}(t)\cos\eta(t)\;,\nonumber\\ f_1(t)&:=&i\sqrt{\frac{c_1(t)}{b_1(t)}}=if_0e^{-i\gamma_1(t)}\;,~~~ f_{1,0}:=f_1(0)=if_0\;,\nonumber\\ \eta_1(t)&:=&2\int_0^tE^{(1)}(s)\sin\eta(s)ds\;. \label{eta1} \end{eqnarray} Clearly we can repeat this procedure indefinitely and construct an infinite product expansion for the evolution operator. Again if we compute only a finite number of terms in this expansion, then we obtain a generalization of the adiabatic approximation. The validity of this approximation may be checked by computing the transformed Hamiltonians. It is not difficult to show that the transformed Hamiltonian obtained after $\ell$ adiabatic canonical transformations is of the form \[H_\ell^{(1)}(t)= h_\ell(t) S(t)\;,\] where \begin{eqnarray} h_\ell(t)&=&E^{(1)}(t)\cos\eta(t)\cos\eta_1(t)\cos\eta_2(t) \cdots\cos\eta_{\ell-1}(t)\;,~~~{\rm and}\nonumber\\ \eta_j(t)&:=&2\int_0^t E^{(1)}(s)\cos\eta(s)\cos\eta_1(s)\cdots \cos\eta_{j-1}(s)\sin\eta_{j-1}(s)ds\;,\nonumber \end{eqnarray} where $j\in\{2,3,\cdots,\ell-1\}$ and $S(t)$ is a two-by-two matrix of unit determinant. Clearly if for some $\ell$, $h_\ell(t)$ is negligible, then the above mentioned generalization of the adiabatic approximation is valid. Finally let us note that in general the initial Hamiltonian (\ref{2-level-H-traceless}) can be written in the form \begin{equation} H(t)=\alpha_1(t)\sigma_1+\alpha_2(t)\sigma_2+a(t)\sigma_3\;, \label{linear-H} \end{equation} with $\alpha_1=(b+c)/2$ and $\alpha_2=i(b-c)/2$. Performing the canonical transformation (\ref{H-trans}) defined by $g(t)=\exp\{i\int_0^t \alpha_2(s)ds\sigma_2\}$, we transform the Hamiltonian (\ref{linear-H}) into \begin{eqnarray} H'(t)&=&\alpha'(t)\sigma_1+a'(t)\sigma_3=\left(\begin{array}{cc} a'(t) & \alpha'(t)\\ \alpha'(t) & -a'(t)\end{array}\right)\;,~~~{\rm where} \label{linear-H'}\\ \alpha'(t)&:=&\alpha_1(t)\cos\xi(t)-a(t)\sin\xi(t)\;,~~~~ a'(t):=\alpha_1(t)\sin\xi(t)+a(t)\cos\xi(t)\;,\nonumber\\ \xi(t)&:=&\int_0^t\alpha_2(s)ds\;.\nonumber \end{eqnarray} Here we have used Eqs.~(\ref{H-trans}) and (\ref{identity}). Next we perform another canonical transformation, namely the one defined by $g(t) =\exp\{i\int_0^t a'(s)ds\sigma_3\}$. This transformation maps the Hamiltonian (\ref{linear-H'}) into \begin{equation} H''(t)=\alpha'(t) e^{i\eta'(t)\sigma_3/2} \sigma_1 e^{-i\eta'(t)\sigma_3/2}= \alpha'(t) [\cos\eta'(t)\sigma_1-\sin\eta'(t)\sigma_2]=\alpha'(t) \left(\begin{array}{cc} o& e^{i\eta'(t)}\\ e^{-i\eta'(t)} & 0\end{array}\right)\;, \label{linear-H''} \end{equation} where $\eta'(t):=2\int_0^t a'(s)ds$. This Hamiltonian is not only a member of Class~3 Hamiltonians, but initially (at $t=0$) its off-diagonal matrix elements are equal. In particular, it has the properties 1. in the above list. Note that we can carry out these canonical transformations on any two-level Hamiltonian. Therefore, every two-level Hamiltonian is canonically equivalent to a Class~3 Hamiltonian of the form~(\ref{linear-H''}). This means that the results obtained for Class~3 Hamiltonians apply to arbitrary two-level Hamiltonians. \section*{V. Time-dependent Simple Harmonic Oscillator} It is well-known that the solution of every second order linear ODE \cite{ode} can be reduced to the classical equation of motion for a simple harmonic oscillator with a time-dependent frequency $\omega=\omega(t)$, \begin{equation} \ddot x(t)+\omega^2(t) x(t)=0\;. \label{sho} \end{equation} It is also well-known that one can reduce both the classical and quantum equations of motion for a generalized harmonic oscillator to Eq.~(\ref{sho}), \cite{hoso,manko,jackiw,jpa98-1}. This equation has, therefore, many physical applications \cite{manko,application}. Yet an exact analytic expression for the general solution of this equation is not known even for the case of real frequency \cite{belman}. \footnote{The lack of an exact analytic solution of Eq.~(\ref{sho}) is not surprising. One way to see this is to recall that the time-independent Schr\"odinger equation for an arbitrary potential $V(x)$ in one dimension is given by \begin{equation} \frac{d^2\psi_n}{dx^2}+\left(\frac{\hbar^2[E_n-V(x)]}{2m}\right)\psi_n=0, \label{time-indep} \end{equation} where $E_n$ and $\psi_n$ are the energy eigenvalues and eigenfunctions, respectively. Eq.~(\ref{time-indep}) can be easily identified with Eq.~(\ref{sho}) provided that one makes the change of variables: $x\to t,~\psi_n\to x$, and $\{\hbar^2[E-V(x)]\}/(2m)\to \omega^2(t)$. This shows that if one was able to find the exact analytic solution of Eq.~(\ref{sho}) for arbitrary frequency $\omega$, then one would have been able to find the general solution of the time-independent Schr\"odinger equation for any potential $V$.} In the following we shall consider the case of an ordinary time-dependent harmonic oscillator (\ref{sho}) with real frequency. In order to apply the results of the preceding section to Eq.~(\ref{sho}), we first express it in the form of a system of first order ODEs. Defining, \[ |\psi(t)\rangle:=\left(\begin{array}{c} x(t)\\ v(t)\end{array}\right)\;,~~~ {\rm and}~~~v(t):=\dot x(t)\;,\] we can write Eq.~(\ref{sho}) in the form of the Schr\"odinger equation (\ref{sch-eq}) with a two-level Hamiltonian of the form (\ref{2-level-H-traceless}) with \begin{equation} a=0\;,~~~~b=i\;,~~~~c=-i\omega(t)^2\;,~~{\rm and}~~E=\omega(t)\;. \label{abc-sho} \end{equation} Since $a=0$, this system belongs to the Class~3 of the preceding section with \begin{eqnarray} f(t)&=&\omega(t)\;,~~~\eta(t)=2\int_0^t\omega(s)ds\;,\nonumber\\ H^{(1)}&=&E^{(1)}(t)\left(\begin{array}{cc} \cos\eta(t)&\frac{\sin\eta(t)}{\omega_0}\\ \omega_0\sin\eta(t)&\cos\eta(t)\end{array}\right)\;,~~~ {\rm and}~~~E^{(1)}(t):=\frac{i\dot\omega(t)}{2\omega(t)}. \label{H1-sho} \end{eqnarray} Clearly, for real frequency $\omega(t)$ we can scale the time variable $t$ so that $\omega_0=1$. Then the transformed Hamiltonian (\ref{H1-sho}) takes the form \begin{equation} H^{(1)}(t)=E^{(1)}(t) \left[\sin\eta(t)\sigma_1+\cos\eta(t)\sigma_3 \right]=E^{(1)}(t) e^{-i\eta(t)\sigma_2/2}\sigma_3 e^{i\eta(t)\sigma_2/2}\;. \label{H1-sho-0} \end{equation} Note that since $\eta(t)$ is real, the Hamiltonian (\ref{H1-sho-0}) is an anti-Hermitian matrix with orthogonal eigenvectors. Therefore, up to a factor of $i$ it describes a two-level spin system with a $SU(2)$ dynamical group \cite{su2,jmp97-2}.\footnote{Note that one can absorb the factor of $i$ in the definition of the time variable $t$, i.e., by defining the imaginary time variable $\tau:=-it$. Therefore, the dynamics given by the Hamiltonian (\ref{H1-sho-0}) may be viewed as the dynamics of a spin system with imaginary time.} This is rather surprising, for it is well-known that the quantum harmonic oscillator has a $SU(1,1)$ dynamical group and that its Schr\"odinger equation may be reduced to Eq.~(\ref{sho}) by means of a quantum canonical transformation corresponding to a time-dependent dilatation \cite{jpa98-1}. In view of the fact that $E^{(1)}$ is proportional to the derivative of $\ln\omega$, we can make a change of independent variable, namely $t\to\eta$. Note that $\eta$ is the integral of a positive real function of $t$. Hence, it is a monotonically increasing function of $t$. Making this change of variable the Schr\"odinger equation for the Hamiltonian (\ref{H1-sho-0}) becomes \[i\frac{d}{d\eta}\tilde U(\eta)=\tilde H(\eta)\tilde U(\eta)\;,~~~ \tilde U(\eta)=1,\] where $\tilde U(\eta):=U'(t(\eta))$, $U'(t)$ is the evolution operator for the the Hamiltonian (\ref{H1-sho-0}), \begin{eqnarray} \tilde H(\eta)&:=&\tilde E(\eta) e^{-i\eta\sigma_2/2}\sigma_3 e^{i\eta\sigma_2/2}=\tilde E(\eta)(\sin\eta\sigma_1+\cos\eta\sigma_3)\;, \label{tilde-h}\\ \tilde E(\eta)&:=&\frac{i\omega'(\eta)}{2\omega(\eta)}\;,~~{\rm and}~~ \omega':=\frac{d\omega}{d\eta}\;. \label{E-eta} \end{eqnarray} Up to a factor of $i$, the Hamiltonian (\ref{tilde-h}) describes the interaction of a spin 1/2 magnetic dipole with a changing magnetic field whose direction rotates uniformly in the $x-z$ plane. As we mentioned in the preceding section for the Class 3 systems $H^{(2)}(t)=H(t)$. Hence direct application of the method of the adiabatic product expansion does not lead to a solution of the Schr\"odinger equation for the Hamiltonian (\ref{H1-sho-0}) or (\ref{tilde-h}). In this case, either one constructs the modified adiabatic product expansion of the preceding section or examines the adiabatic series expansion of Ref.~\cite{pra97-1}. The latter yields a series expansion for the evolution operator $\tilde U(\eta)$ of the Hamiltonian $\tilde H(\eta)$, namely \begin{eqnarray} \tilde U(\eta)={\cal T}e^{-i\int_0^\eta \tilde H(s)ds}&=& 1-i\int_0^\eta \tilde H(s)ds+\frac{(-i)^2}{2} \int_0^\eta\int_0^\eta {\cal T}[\tilde H(s_1)\tilde H(s_2)]ds_1ds_2+\cdots+\nonumber\\ &&\hspace{2mm}\frac{(-i)^n}{n!}\int_0^\eta\cdots\int_0^\eta {\cal T}[\tilde H(s_1) \cdots \tilde H(s_n)] ds_1\cdots ds_n +\cdots\;, \label{dyson} \end{eqnarray} where ${\cal T}$ stands for the time-ordering operator. Since $\tilde H(\eta)$ is proportional to $\omega'(\eta)$, for slowly varying $\omega$ one obtains an approximate expression for $\tilde U(\eta)$ by computing a finite number of terms in this series. This is in fact another generalization of the adiabatic approximation, because if one keeps only the first term in this series and neglects the other terms one is essentially neglecting $\tilde H$ or alternatively $H^{(1)}$. As we explained above, this is just the adiabatic approximation. If one keeps more terms in this series, then one obtains a better approximation than the adiabatic approximation. \section*{VI. Quadrupole Interaction of a Spin 1 Particle with a Changing Electric Field} Consider a spin 1 particle interacting with a changing electric field $\vec{\cal E}(t)=({\cal E}_1(t),{\cal E}_2(t),{\cal E}_3(t))$ according to the Stark Hamiltonian \begin{equation} H(t)=\lambda[\vec J\cdot\vec{\cal E}(t)]^2\;, \label{stark} \end{equation} where $\lambda$ is a real coupling constant and $\vec J$ is the angular momentum of the particle. The quadrupole interactions of the form (\ref{stark}) have been extensively studied for fermionic systems in relation with the non-Abelian geometric phases \cite{mead,avron1,avron2} (See also \cite{j-a}.) The occurrence of non-Abelian geometric phases for the degenerate spin ~1 systems has been pointed out in Ref.~\cite{jpa97}. For these systems, the particle has a definite angular momentum $j=1$ and the Hamiltonian is a $3\times 3$ matrix. Using the spin $j=1$ representation of $J_i$, we can express the Stark Hamiltonin (\ref{stark}) in the form \begin{equation} H=(\frac{\lambda r^2}{2}) \left(\begin{array}{ccc} 1+2z^2 & \sqrt{2} z e^{-i\theta} & e^{-2i\theta} \\ \sqrt{2} z e^{i\theta} & 2 & -\sqrt{2} z e^{-i\theta} \\ e^{2i\theta} &-\sqrt{2} z e^{i\theta}& 1+2z^2 \end{array}\right)\;, \label{stark-matrix} \end{equation} where $r,~\theta,$ and $z$ are defined by \[r:=\sqrt{{\cal E}_1^2+{\cal E}_2^2}\;,~~~e^{i\theta}:= \frac{{\cal E}_1+i{\cal E}_2}{r}\;,~~ {\rm and}~~z:=\frac{{\cal E}_3}{r}\;.\] In view of the general results of Ref.~\cite{jpa97}, if $r\neq 0$ then the Hamiltonian (\ref{stark-matrix}) has a degenerate and a nondegenerate eigenvalue. In the following we shall consider the case where ${\cal E}_3=0$. The general case ${\cal E}_3\neq 0$ can be similarly treated. If ${\cal E}_3=0$, then $z=0$ and \begin{equation} H=(\frac{\lambda r^2}{2}) \left(\begin{array}{ccc} 1 & 0 & e^{-2i\theta} \\ 0 & 2 & 0 \\ e^{2i\theta} & 0 & 1 \end{array}\right)\;. \label{stark-matrix-0} \end{equation} The eigenvalues of this Hamiltonian are given by \begin{equation} E_1=0,~~~~E_2=\lambda r^2\;. \label{S-eg-va} \end{equation} For $r\neq 0$, $E_1$ is nondegenerate and $E_2$ is doubly degenerate. A set of orthonormal eigenvectors of this Hamiltonian is given by \begin{equation} |\psi_1;R\rangle:=\frac{1}{\sqrt{2}}\left(\begin{array}{c} -1\\ 0\\e^{2i\theta}\end{array}\right)\;,~~~ |\psi_2,1;R\rangle:=\frac{1}{\sqrt{2}}\left(\begin{array}{c} 1\\ 0\\e^{2i\theta}\end{array}\right)\;,~~~ |\psi_2,2;R\rangle:=\left(\begin{array}{c} 0\\ 1\\0\end{array}\right)\;. \label{S-eg-ve} \end{equation} where $R=(r,\theta)$. Next we compute $U^{(0)}(t)$ for this system. In order to do this we first use Eq.~(\ref{a=}) to calculate ${\cal A}^n$. In view of the fact that the Hamiltonian (\ref{stark-matrix-0}) is Hermitian, $|\phi_n,a;R\rangle=|\psi_n,a;R\rangle$ and Eq.~(\ref{a=}) leads to \begin{equation} {\cal A}^1=-d\theta\;,~~~{\cal A}^2=\left(\begin{array}{cc} -d\theta&0\\ 0&0\end{array}\right)\;. \label{S-a=} \end{equation} Substituting Eqs.~(\ref{S-a=}) into Eq.~(\ref{6}) and making use of Eq.~(\ref{S-eg-va}), we find \begin{equation} K^1(t)=e^{-i[\theta(t)-\theta_0]} \;,~~~~ K^2(t)=e^{-i\rho(t)}\left(\begin{array}{cc} e^{-i[\theta(t)-\theta_0]}&0\\ 0&1\end{array}\right)\;, \label{K's} \end{equation} where \begin{equation} \theta_0:=\theta(0)\;,~~{\rm and}~~\rho(t):= \lambda\int_0^t r(s)^2ds\;. \label{theta-rho} \end{equation} Using Eqs.~(\ref{u0}) and (\ref{S-eg-ve}), we have \begin{equation} U^{(0)}(t)=\frac{1}{2}\left(\begin{array}{ccc} (1+e^{-i\rho(t)})e^{i\theta_-(t)}&0&(-1+e^{-i\rho(t)})e^{-i\theta_+(t)}\\ 0 & 2e^{-i\rho(t)} & 0\\ (-1+e^{-i\rho(t)})e^{i\theta_+(t)}&0&(1+e^{-i\rho(t)})e^{i\theta_-(t)} \end{array}\right)\;, \label{S-u0} \end{equation} where $\theta_\pm(t):=\theta(t)\pm\theta_0$. Next we compute the Hamiltonian $H^{(1)}(t)$. This involves the calculation of $A_{ab}^{nm}(t)$ and $H_{ab}^{mn}(t)$ for $m\neq n$. Using Eqs.~(\ref{H1-ab}) and (\ref{S-eg-ve}) we have \begin{eqnarray} A^{12}_{11}&=&A^{21}_{11}=-\dot\theta\;,~~~ A^{12}_{12}=A^{21}_{21}=0\;,\nonumber\\ H^{21}_{11}&=&H^{12}_{11}=\dot\theta e^{i\rho(t)}\;,~~~ H^{12}_{12}=H^{21}_{21}=0\;.\nonumber \end{eqnarray} Substituting these equations in Eq.~(\ref{H1}) and using Eqs.~(\ref{S-eg-ve}), we obtain \begin{equation} H^{(1)}(t)=-\dot\theta(t) \left(\begin{array}{ccc} \cos\rho(t)&0&-i\sin\rho(t) \\ 0& 0& 0\\ i\sin\rho(t) &0&-\cos\rho(t) \end{array}\right)=-\dot\theta(t)[\sin\rho(t)\Sigma_2+ \cos\rho(t)\Sigma_3]\;, \label{S-H1} \end{equation} where \begin{equation} \Sigma_2:=\left(\begin{array}{ccc} 0&0&-i\\ 0&0&0\\ i&0&0\end{array}\right) \;,~~{\rm and}~~ \Sigma_3:=\left(\begin{array}{ccc} 1&0&0\\ 0&0&0\\ 0&0&-1\end{array}\right)\;. \label{sigma23} \end{equation} It is not difficult to recognize $\Sigma_2$ and $\Sigma_3$ as the Pauli matrices $\sigma_2$ and $\sigma_3$ represented in a $(0+1/2)$ representation of $SU(2)$. In view of this identification we can express $H^{(1)}(t)$ in the form \begin{equation} H^{(1)}(t)=-\dot\theta(t) e^{i\rho(t)\Sigma_1/2}\Sigma_3\; e^{-i\rho(t)\Sigma_1/2}\;, \label{S-H1-su2} \end{equation} where \[\Sigma_1:=\left(\begin{array}{ccc} 0&0&1\\ 0&0&0\\ 1&0&0\end{array}\right)\;,\] and we have used Eq.~(\ref{identity}). The Hamiltonian $H^{(1)}(t)$ has the following interesting properties. \begin{itemize} \item[a)] For $\dot\theta=0$, i.e., $\theta=$ constant, the adiabatic approximation is exact and $U(t)=U^{(0)}(t)$. \item[b)] In view of Eq.~(\ref{S-H1-su2}) for $\dot\theta\neq 0$, $H^{(1)}(t)$ has three nondegenerate eigenvalues, namely $-\dot\theta,~0,$ and $\dot\theta$. This is quite remarkable because it shows that the adiabatic canonical transformation maps the degenerate Hamiltonian (\ref{stark-matrix-0}) into the nondegenerate Hamiltonian (\ref{S-H1-su2}). \item[c)] Since $\Sigma_i$ is a representation of the Pauli matrix $\sigma_i$, the Hamiltonian (\ref{S-H1-su2}) belongs to a representation of the Lie algebra of $SU(2)$. This means that one can reduce the Schr\"odinger equation for this Hamiltonian to that of the dipole Hamiltonian \cite{pra97-1,jmp97-2} \[H_{\rm dp}=-2\dot\theta(t) e^{i\rho(t)J_1}J_3e^{-i\rho(t)J_1}\;.\] \item[d)] We can perform another canonical transformation, namely the one defined by $g(t)=\exp[-i\rho(t)\Sigma_1/2]$ to transform the Hamiltonian (\ref{S-H1-su2}) into \begin{equation} H^{(1)'}(t)=\frac{r(t)^2}{2}\Sigma_1-\dot\theta(t)\Sigma_3\;, \label{S-H1-prime} \end{equation} where we have used Eqs.~(\ref{S-H1-su2}), (\ref{H-trans}) and (\ref{theta-rho}). In particular if $\dot\theta$ and $r^2$ happen to be proportional, i.e., for some $c\in\relax{\rm I\kern-.18em R}$ \begin{equation} \dot\theta(t)=c\;r(t)^2\;, \label{condi-ex} \end{equation} then $H^{(1)'}(t) =r(t)^2(\frac{1}{2}\Sigma_1-c\Sigma_3)$. In this case the eigenvectors of $H^{(1)'}(t)$ are constant and the adiabatic approximation yields the exact solution of the Schr\"odinger equation for $H^{(1)'}(t)$. The corresponding evolution operator is then given by \begin{equation} U^{(1)'}(t) =e^{-i(\frac{1}{2}\Sigma_1-c\Sigma_3)\int_0^t r(s)^2ds}\;. \label{u1-prime} \end{equation} Having obtained the evolution operator for $H^{(1)'}(t)$ we can use Eq.~(\ref{u-trans}) to obtain the evolution operator for $H^{(1)}(t)$ and $H(t)$. This yields the following expression for the evolution operator for $H(t)$: \begin{equation} U(t)=U^{(0)}(t) e^{i\rho(t)\Sigma_1/2} U^{(1)'}(t) \;, \label{u=ex} \end{equation} where $U^{(0)}(t)$ and $U^{(1)'}(t)$ are given by Eqs.~(\ref{S-u0}) and (\ref{u1-prime}). \end{itemize} The above analysis shows that the condition~(\ref{condi-ex}) defines a class of exactly solvable time-dependent Stark Hamiltonians. If $\theta=\omega t$, for some constant frequency $\omega$, this condition corresponds to the case of the rotating electric field $\vec E=r( \sin\omega t,\cos\omega t,0)$ with magnitude $r$. \section*{VII. Conclusion} In this article we have extended the method of the adiabatic product expansion to non-Hermitian and degenerate Hamiltonians. We showed that in general there were three possibilities for the adiabatic product expansion: \begin{itemize} \item[1)] The expansion terminates after a finite number of iterations. This happens when one of the transformed Hamiltonians vanishes. In this case the method yields the exact solution for the Schr\"odinger equation; \item[2)] The expansion consists of an infinite number of distinct terms. In this case, the method does not lead to an exact solution, but it gives rise to a generalization of the adiabatic approximation. This approximation is performed by keeping a finite number of terms in the product expansion. The general asymptotic behaviour of the adiabatic product expansion has not been studied. However, one can interpret this approximation by recalling that the condition for the termination of the product expansion corresponds to the validity of the conventional adiabatic approximation for one of the transformed Hamiltonians. \item[3)] The expansion involves terms which are not distinct. In this case the expansion does not lead to a solution. However, usually one can make another time-dependent canonical transformation after each adiabatic transformation and obtain an infinite product expansion with the properties of case 2) above. \end{itemize} We have considered some specific problems that one can attempt to solve using this method. We treated the case of a general nondegenerate two-level system and applied our general results to the more specific case of the classical equation of motion for a harmonic oscillator with a time-dependent frequency. In this case, we showed that the adiabatic canonical transformation mapped the corresponding two-level quantum system to a quantum system with an anti-Hermitian Hamiltonian. Although the direct application of the method of adiabatic product expansion did not yield a solution, we could construct the modified adiabatic product expansion. We have also outlined an adiabatic series expansion for the time-evolution operator of this system which led to another generalization of the adiabatic approximation. Finally, we considered the application of our method to treat the quadrupole interaction of a spin~1 particle with a changing electric field. The corresponding (Stark) Hamiltonian had a nondegenerate as well as a degerenrate eigenvalue. We showed that the adiabatic canonical transformation mapped this Hamiltonian to a Hamiltonian which had nondegenerate eigenvalues and belonged to a reducible $(0+1/2)$ representation of the Lie algebra of $SU(2)$. This means that we can directly use the results of Refs.~\cite{pra97-1,jmp97-2} which treat the Schr\"odinger equation for a nondegenerate Hamiltonian belonging to (an irreducible representation of) the Lie algebra of $SU(2)$. Furthermore, we identified a class of exactly solvable spin 1 quadruple Hamiltonians.
\section{Introduction} It is well known that most of the astronomical informations are carried by massless spinning particles. Astronomers usually extract astronomical informations from photons, which are spin-1 massless particles. In 1987 astronomers succeeded in extracting informations carried by nutrinos (sipn- one half massless particles) coming from SN1987A. In the near future we expect gravitons (spin-2 massless particles) to open a third window through which we look at the universe. These particles, during their trip from their sources to the detectors, are moving in different gravitational fields. Thus, it is of fundamental importance, for astronomy and space studies, to know the exact trajectories of such particles, and the interaction between the spin of these particles and the background gravitational field. The problem of motion of spinning \footnote{In the present work we are going to distinguish between spin and rotation. We use the term spin for the quantum property of a microscopic particle, while the term rotation is used for the corresponding classical mechanical property of a macroscopic object.} \ particles in a background gravitational field has been tackled by different authors. The trails of those authors can be classified into two main approaches. {\it The first approach} contains two classes: (i) quantization of a relativistic equation of motion of a rotating object (e.g.Papapetrou equation, Dixon equation),(cf. Melek (1988)) (ii) geometrization (or more specifically coveriantization) of a quantum mechanical equation of motion of a spinning particle (e.g. Schr\"{o}denger equation, Pauli equation),(cf. DeWitt (1957), Galvao and Teitelboim (1980)). {\it The second approach} depends on a different philosophy in which paths (or curves) of an appropriate geometry are considerd to represent trajectories of test particles . Einstein had followed this approach and used the geodesic and null geodesic curves of the Riemannian geometry to describe the motion of test particles and of photons respectively. Although the philosophy of the second approach is successful in describing motion of test particles including massless particles (photons), yet it was neglected by many authors in dealing with motion of other spinning particles. We believe that this philosophy deserves further investigations especially to look for the equation of motion of spinning particles in gravitational fields. It is to be noted that while the first approach is suitable for describing short range motion of spinning particles i.e. motion on the laboratory scale, it is not suitable for describing long range motion (e.g. motion on scales such as solar system, galactic, intergalactic scales). The second approach is more suitable for describing trajectories of test particles especially long range motion of massless particles in gravitational fields (motion of photons using null-geodesics). We are mainly interested in this approach. To explore the capabilities of this approach we directed our attention to the AP-geometry. The cause is that short range motion of spinning particles is described successfully using this geometry, since spinors can be defined and related to the structure of the AP-spaces. In a trial to look for possible paths in this geometry, three path equations were derived (Wanas et. al. (1995)a) by generalizing the method given by Bazanski (1977,1989). These equations can be written in the form, $$ {\frac{dV^\mu}{dS^+}} + \{^{\mu}_{\alpha\beta}\} V^\alpha V^\beta = - \Lambda^{~ ~ ~ ~ \mu}_{(\alpha \beta) .} ~~~V^\alpha V^\beta, \eqno{(1.1)} $$ $$ {\frac{dW^\mu}{dS^0}} + \{^{\mu}_{\alpha\beta}\} W^\alpha W^\beta = - {\frac{1}{2}} \Lambda^{~ ~ ~ ~ \mu}_{(\alpha \beta) .}~~~ W^\alpha W^\beta, \eqno{(1.2)} $$ $$ {\frac{dU^\mu}{dS^-}} + \{^{\mu}_{\alpha\beta}\} U^\alpha U^\beta = 0, \eqno{(1.3)} $$ where$S^{+}, S^{0}$ and $S^{-}$ are the evolution parameters characterizing the three paths respectively ; and $V^{\alpha}, W^{\alpha}$ and$U^{\alpha}$ are the tangents to the corresponding paths and the brakets () are used for symmetrization. The torsion of the AP-space is defiend by $$ \Lambda^\alpha_{. \mu \nu}{\ } {\stackrel{def.}{=}} {\ } \Gamma^\alpha_{. \mu \nu} - \Gamma^\alpha_{. \nu \mu} {\ }{\ }, $$ where $\Gamma^{\alpha}_{.\mu\nu}$ is a non-symmetric affine connexion defiend as a consequance of the AP-condition (see section.2). These equations can be considered as generalization of the geodesic equation of Riemannian geometry. Considering these three equations, it seems interesting to point out the following remarks: 1- Although four absolute derivatives are defiend in the AP-space (see section 2), and used in deriving the above equations, only three equations were obtained inculding the geodesic equation (1.3). 2- In general, the other two equation ((1.1),(1.2)) can not be reduced to the geodesic equation (1.3) unless the torsion vanishes . It has been shown (Wanas,and Melek (1995) ) that the vanishing of the torsion of the AP-space will lead to the vanishing of the curvature tensors defined in that space. In this case the space will reduce to a flat one. 3- Equation (1.3) can be reduced to null-geodesic upon reparametrization. 4- The equations can be considered as geodesic equations modified by a torsion term. The important feature of this set of equations is that the numerical coefficient of the torsion term jumps by a step of one half from one equation to the next. This is tempting to beleive that paths in the AP- spaces are quantized in a certain sense. From these remarks it seems that the AP-space possesses more finer structure than that appearing in its conventional picture (for more details about the conventional sturture of the AP-spaces, (cf. Levi-Civita (1950), Mikhail (1952), Hayashi and Shirafuji (1979)). It contains in addition to the geodesic and null- geodesic equations, two more equations (1.1), (1.2). Now the following groups of questions emerge: 1- Does the space possesses a general affine connexion that gives rise to family of paths in which the coefficient of the torsion term jumps by a step of $\frac{1}{2}$ from equation to another in this family ? If so, what is the underlying geometric structure ? 2- What is the general form of the equations representing this family with a jumping torsion term ? 3- Is there any observational or experimental evidence for motion alonge such paths ? What is the order of magnitude of the deviation from the geodesic motion, if any? What is the intrinsic property of the moving particle that causes such deviation from the geodesic one? The aim of the present work is to give answers to some of the above questions. In section 2, a trail to answer the first group of the above questions is given. In section 3 the general form of a path equation, giving the required family mentioned above, is derived using the geometry given in section 2. Section 4 contains a trail to attribute some physical meaning to the new path equation. The static weak field limits of the new equation are given in section 5. The work is discussed in section 6. \section{The Underlying Geometry} In the conventional AP-geometry two affine connexions were defined. The first is Christoffel symbol defined as a consequence of the metricity condition, $$ g_{\alpha \beta ; \mu} = 0, \eqno{(2.1)} $$ where $g_{\alpha \beta}$ is the metric tensor defiend by, $$ g_{\alpha \beta} {\ } {\stackrel{def.}{=}} {\ } \h{i}_{\alpha} \h{i}_{\beta} {\ }{\ }{\ }, \eqno{(2.2)} $$ and $\h{i}^{\alpha}$ are the tetrad vectors defining the structure of the AP-space in 4-dimensions. The semicolon is used to characterize covarient dervatives using Christoffel symbols. The second is the non-symmetric connexion ${\Gamma}^{\mu}_{.\alpha \beta}$ defiend as a consequence of absolute parallelism condition: $$ {{\h{i}}^{{\stackrel{\alpha}{+}}}_{~~.| \beta}} = 0 {\ }{\ }{\ }, \eqno{(2.3)} $$ the stroke and the (+) sign is used to characterize absolute derivatives using ${\Gamma}^{\mu}_{.\alpha \beta}$. Since this connexion is non-symmetric, one can define two more absolute derivatives : for the first we use a stroke and a (-) sign which indicates the use of the dual connexion ${\hat{\Gamma}}^{\mu}_{.\alpha \beta} (= {\Gamma}^{\mu}_{.\beta \alpha})$, and for the second we use a stroke without signs, which indicates the use of the symmetric part, $\Gamma^{\mu}_{.(\alpha \beta)}$. It can be shown that using (2.2), (2.3) we get, $$ g_{{\stackrel{\mu}{+}}{\stackrel{\nu}{+}}| \sigma} =0 {\ }{\ }{\ }, \eqno{(2.4)} $$ which means that metricity is also preserved using the non-symmetric connexion, recalling that the metricity condition (2.1) is necessary but not sufficient to define Cristoffel symbol. Now we are looking for a general affine connexion that is capable of producing the family of all paths with jumping torison term .Recalling that the three path equations were derived using the above mentioned derivatives ,thus the simplest way to find such a connexion is to combine linearly the above connexion using some parameters.After some manipulations the general expression of this connexion can be written in the following form, $$ \nabla^\mu_{.\alpha \beta} = a_1 \{^\mu_{\alpha \beta}\} + (a_2 - a_3) \Gamma^{\mu}_{. \alpha \beta} - (a_3 + a_4)\Lambda^{\mu}_{. \alpha \beta}{\ }{\ }{\ }, \eqno{(2.5)} $$ where $a_1 , a_2, a_3$ and $a_4$ are parameters to be fixed later. It can be easily shown that $\nabla^{\alpha}_{.\mu \nu}$ transforms as an affine connexion under the group of general coordinate transformations. It is clear that this connexion is non-symmetric. If we charactize the general absolute derivative, using the affine connexion (2.5), by a double stroke we get after some manipulations: $$ g_{\mu \nu || \sigma} = (1- a_1 - a_2 - a_3 -2a_4)(g_{\mu \alpha} \{^{\alpha}_{\sigma\nu}\} + g_{\nu \alpha} \{^{\alpha}_{\mu\sigma}\}) - (a_3 + a_4)(g_{\mu \alpha}\gamma ^{\alpha}_{. \sigma \nu} + g_{\nu \alpha}\gamma^{\alpha}_{. \sigma\mu}){\ }, \eqno{(2.6)} $$ where ${\gamma}^{\alpha}_ {.\mu \nu} $ is the contorsion defined by $$ \gamma^{\alpha}_{. \mu \nu} {\ } {\stackrel{def.}{=}} {\ } \frac{1}{2}(\Lambda^{\alpha}_{.\mu \nu} - {\Lambda}_{\mu .\nu }^{~\alpha} - {\Lambda}_{\nu .\mu}^{~\alpha}) \eqno{(2.7)} $$ If we need metricity to be prserved along the paths characterized by (2.5) we should take, $$ a_3+a_4=0. $$ $$ 1 - a_1 - a_2 + a_3 = 0, $$ Taking $a=a_1$, $b=a_2+a_4$ then the metricity condition can be written in the form $$ a+b=1 . \eqno{(2.8)} $$ After using this condition it is clear that (2.5) will remain non-symmetric. Using the general affine connexion (2.5) and the general metricity condition (2.8), one can define the following curvature tensors. (i) If we replace, in the definition of Riemann-Christoffel tensor $R^{\alpha}_{.\mu \nu \sigma}$, Christoffel symbol by the new connexion we get the curvature, $$ B^{\alpha}_{.\mu \nu \sigma} {\ } {\stackrel{def.}{=}} {\ } \nabla^{\alpha}_{.\mu \sigma ,\nu} - \nabla^{\alpha}_{.\mu \nu ,\sigma} + \nabla^{\alpha}_{.\epsilon \nu} \nabla^{\epsilon}_{.\mu \sigma } - \nabla^{\alpha}_{.\epsilon \sigma } \nabla^{\epsilon}_{.\mu \nu }{\ }{\ }{\ }, \eqno{(2.9)a} $$ which can be written in the form $$ B^{\alpha}_{.\mu \nu \sigma} = R^{\alpha}_{.\mu \nu \sigma} + b {\gamma}^{\alpha}_ {.\mu [\sigma ; \nu]} +b^2 {\gamma}^{\alpha}_ {.\epsilon [\nu} {\gamma}^{\epsilon}_ {.\mu \sigma ]} {\ }{\ }{\ }, \eqno{(2.9)b} $$ the square brackets are used for antisymmetrization. (ii) If we define the curvature $W^\alpha_{.\mu \nu \sigma}$ as a measure of the non-commutation of the general absolute derivatives, we get $$ \h{i}_{\mu || \nu \sigma} - \h{i}_{\mu || \sigma \nu} = \h{i}_{\alpha}{\ }W^{\alpha}_{. \mu \nu \sigma} \eqno{(2.10)} $$ where $$ W^\alpha_{.\mu \nu \sigma} = B^\alpha _{.\mu \nu \sigma} - b(b-1) \gamma ^\alpha_{.\mu \epsilon} \Lambda^\epsilon_{.\nu \sigma} \eqno{(2.11)} $$ Now we have the following remarks: (1) It is to be considered that non of the curvature tensors defined by (2.9) and (2.11) vanishes. This means that the new structure of the AP-space is, in general not flat. This result will be discussed in section 6. (2) Taking the metricity condition into consideration, we can show that two important special cases could be obtained: {\it The first case} a=1 (i.e. b=0): In this case the geometry will be reduced to Riemannian geometry. This is obvious by substituting the values of the parameters into (2.5), (2.9)b, and (2.11) , which give , $$ \nabla^\alpha _{.\mu \nu} = \{^{\alpha}_{\mu \nu}\} \eqno{(2.12)} $$ $$ W^\alpha _{.\mu \nu \sigma} = B^\alpha_{.\mu \nu \sigma} = R^\alpha _{.\mu \nu \sigma} \eqno{(2.13)} $$ {\it The second case} a=0 (i.e. b=1): In this case it can be easily shown, after some manipulations, that the geometry reduces to the conventional AP-geometry with $$ \nabla^{\alpha}_{.\mu \nu} = \Gamma^{\alpha}_{.\mu \nu} \eqno{(2.14)} $$ $$ B^{\alpha}_{.\mu \nu \sigma} {\ } = {\ } \Gamma^{\alpha}_{.\mu \sigma, \nu} - \Gamma^{\alpha}_{.\mu \nu, \sigma} + \Gamma^{\alpha}_{\epsilon \nu} \Gamma^ {\epsilon}_{. \mu \sigma} - \Gamma^{\alpha}_{. \epsilon \sigma} \Gamma^{\epsilon}_{. \mu \nu} = 0 \eqno{(2.15)} $$ $$ W^{\alpha}_{.\mu \nu \sigma} = 0. \eqno{(2.16)} $$ The curvature (2. 15) vanishes as a consequence of the AP-condition. These two cases will be discussed in the last section. Now to complete the strcture of the space, one should look for a path equation correspoding to the new affine connexion, which can be considered as a generalization of equations (1.1), (1.2) and (1.3). This is done in the following section. \section{The General Form of the Path Equation} Generalizing the Bazanski (1977,1989) Lagrangian, using the affine connexion (2.5) and the condition (2.8), we get $$ L{\ }{\stackrel{def.}{=}}{\ } \h{i}_{\mu} \h{i}_{\nu} Z^{\mu}{{\frac{\nabla{\chi}^\mu}{ \nabla\tau}}} \eqno{(3.1)} $$ $$ Z^{\mu} {\ } {\stackrel{def.}{=}} {\ } {\frac{dx^\mu}{d\tau}}, \eqno{(3.2)} $$ $$ {\frac{\nabla {\chi}^\mu}{ \nabla\tau}} {\ } {\stackrel{def.}{=}} {\ } \chi^{\mu}_{.|| \alpha}Z^{\alpha} \eqno{(3.3)} $$ $$ \chi^{\mu}_{.|| \alpha} {\ } {\stackrel{def.}{=}} {\ } \chi^{\mu}_{., \alpha} + \chi^{\beta}\nabla^{\mu}_{. \beta \alpha} , \eqno{(3.4)} $$ where $\tau$ is the evolution parameter along the new general path associated with (2.5), $\chi^{\mu}$ is the deviation vector, and $Z^{\mu}$ is the tangent to the path. Applying the variational formalism using the lagrangian (3.1), and noting that the general absolute differentional operation commmutes with rasing and lowering indices (because of (2.8)), we get after necessary manipulation : $$ {{\frac{dZ^\mu}{d\tau}} + \cs{\nu}{\sigma}{\mu}\ Z^\nu Z^\sigma = - b{\ } \Lambda_{(\nu \sigma)}.^\mu~~ Z^\nu Z^\sigma} .\eqno{(3.5)a} $$ Using (3.3), the last equation can be written in the form $$ {\frac{\nabla Z^\mu}{ \nabla\tau}} = 0 . \eqno{(3.5)b} $$ Equation (3.5)a, or (3.5)b, represents a generalization of the path equations given in section 1. Moreover, if $b=0$ this equation will reduce to geodesic of the metric (or null-geodesic upon reparametrization). This is consistent with the first special case given in the above section. It can be easily shown that (3.5) will give rise to, $$ \frac{d(g_{\alpha \beta}Z^\alpha Z^\beta)}{d\tau} = 0 $$ consequently, its first integral is given by, $$ Z^\alpha{\ }Z_{\alpha} = Z^2 \eqno{(3.6)} $$ which means that Z is a constant along the path (3.5), and since Z is scalar under general coordinate transformation, then it will be constant in general. \section{Physical Meaning of the Torsion Term} Equation (3.5) reprsents a new path in the AP-space. As stated above, this equation can be reduced to a geodesic equation (as b=0) which represents the trajectory of a test particle in a background gravitational field (it can also be reduced to a null-geodesic). So, what is the role of the torsion term on the R.H.S. of (3.5)a ? May it repesent a type of interaction between the torsion of the background gravitational field and some intrinsic property of the moving particle ? If so, what is this intrinsic property ? Let us start a trial to answer the last question. Recalling that, the aim of the work in the section 3 is to get an equation that reprsents the complete family of the equations given in section 1. Thus the parameter "b" of (3.5)a should consist of a half integer part. In this case we can write $$ b = {\frac{n}{2}}{\ }\beta , \eqno{(4.1)} $$ where n is a natural number and $\beta$ is another parameter. It is well known that an intrinsic property of the particle which depends on half integers is its quantum spin. Several authors have pointed out that spinning particles feel space-time torsion (cf. Hehl (1971)). Others beleive that spinning particles are the only probes which detect telleparallel geometry(cf. Nitch and Hehl (1980), Ross (1989)). Although most of these authors have used the term spin to mean rotation and have believed that only rotating sources can generate the space-time torsion, the situation here is different. It has been shown by the author(1990) that space-time torsion is generated whether the source of the field is rotating or not. Considering these arguments, we may suggest that the R.H.S. of (3.5)a reprsents a type of interaction between the quantum spin of the moving particle and the torsion of the background gravitational field. Consequently, equation (3.5) may reprsent the trajectory of a spinning particle in a gravitational field. Consequently we take $n = 0, 1, 2,...$ for particles with spin 0 , ${\frac{1}{2}}$, 1,..... respectively. For macroscopic objects and spinless particles $n = 0$.This will reduce equation (3.5) to the geodesic (or null-geodesic) equation, and the geometric structure to a Riemannian one. It is well known that the geodesic motion implies the validity of the weak equivalence principle (WEP). So, equation (3.5) implies that motion of spinning particles violates the WEP. If we take the parameter $\beta$ to be of order unity, then the torsion term will be of the same order of magnitude as the Christoffel symbol term which will be considerably large. But since there are no experimental or observational evidences for such large violation of (WEP) for the motion of spinning particles, thus the parameter $\beta$ should be of less order. From observational point of view the WEP is varified to an accuracy of about $10^{-2}$ on the galactic scale (cf. Longo (1988)), so $\beta$ should be less than $10^{-2}$. From the previous discussion it is acceptable that the parameter $\beta$ may have the following properties : (1) It should be a dimensionless quantity to preserve the dimensions on both sides of (3.5)a. (2) It should be small compared to unity $(\beta \leq 10^{-2})$ to be consistent with relevant obersvations and experiments. (3) It should be connected to the intrinsic properties of elementary particles, especially those affecting motion of the spinning particles. (4) It should include, in its structure, Plank's constant $\hbar$ or h. To the knewledge of the author a quantity satisfying the above requirements is the the fine structure constant $\alpha (={\frac{e^2}{\hbar c}} ={\frac{1}{137}})$. We can replace $\beta$ in (4.1) by the fine structure constant $\alpha$. But to be more conservative we are going to write (4.1) in the form $$ b = {\frac{n}{2}}{\alpha {\ } \gamma} \eqno{(4.2)} $$ where $\gamma$ is a dimensionless parameter of order unity to be fixed by experiment. Now the parameter b constitutes of three parts: the first" ${\frac{n}{2}}$" is the part that makes (3.5)a gives rise to (1.1), (1.2), (1.3), and the second part "$\alpha$" acts as a reduction factor, while the third part"${\gamma}$" is introduced for matching with experimental results. The appearence of the fine structure constant in this treatment will be discussed in the last section. \section{Weak Static Limits of the New Equation} The path equation (3.5) can be used as the equation of motion for any field theory, constructed using the AP-geometry, provided that the theory has good Newotonian limits. In such theories, (e.g. Mikhail and Wanas(1977), M{\o}ller (1978),Hayashi and Shirafuji(1979)), the tetrad vectors $\h{i}_{\mu}$ are considered as field variables. So, if we write $$ \h{i}_{\mu} = \delta_{_i \mu} + \epsilon h_{_i \mu} \eqno{(5.1)} $$ where $\epsilon$ is a small parameter, $\delta_{_i \mu}$ is Kronecker delta and $h_{i \mu}$ reprsents deviations from flat space, then the weak field condition can be fulfilled by neglecting quantities of the second and higher orders in $\epsilon$ in the expanded field quantities. For a static field assumption, we are going to assume the vanishing of time derivatives of the field variables. The vector components $Z^\mu$ defined by (3.2) will have the values $$ Z^1 \approx Z^2 \approx Z^3 \approx \varepsilon {\ }{\ }{\ } , Z^0 \approx 1 - \varepsilon \eqno{(5.2)} $$ where $\varepsilon (\approx \frac{v}{c})$ is a parameter. If we want to add the condition of slowly moving particle to the previous conditions we should neglect quantities of second and higher orders of the parameter $\varepsilon$. Thus, in expanding the quantities of the path equation (4.3) we are going to neglect quantities of orders $\epsilon^2, {\ }\varepsilon^2,{\ } \epsilon \varepsilon$ and higher,and also time derivatives of the field variable are to be neglected. To the first order of the parameters, the only field quantities that will contribute to the path equation (4.3) are given by $$ \Lambda_{00}^{.\ .\ i} = - \epsilon h_{ 0 0, i} ,{\ }{\ }{\ }{\ } (i= 1,2,3) \eqno{(5.3)} $$ $$ \{^{\ i}_{0{\ }0}\} = \frac{\epsilon}{2} Y_{00,i} ,{\ }{\ }{\ }{\ } (i= 1,2,3) \eqno{(5.4)} $$ where $Y_{\mu \nu}$ is defined by $$ g_{\mu \nu} = \eta_{\mu \nu} + \epsilon {\ } Y_{\mu \nu} {\ }{\ }{\ }, $$ $g_{\mu \nu}$ is given by (2.2) and $ \eta_{\mu \nu}$ is the Minkowski metric tensor . Substituting from (5.3),(5.4) into (4.3) we get after some manipulations : $$ \frac{d^2x^i}{d{\tau}^2} = -\frac{1}{2}{\ }\epsilon{\ }(1 -\frac{n}{2}\alpha \gamma) Y_{00,i}{\ }Z^0{\ }Z^0. \eqno{(5.5)} $$ In the present case, the metric of the Riemannian space, associated to AP-space, can be written in the form, $$ (\frac{d\tau}{dt})^2 = c^2{\ }(1 + \epsilon {\ }Y_{00}). \eqno{(5.6)} $$ Substituting from (5.6) into (5.5) we get after some manipulations: $$ \frac{d^2x^i}{dt^2} = - \frac{c^2}{2}{\ }\epsilon{\ }(1 -\frac{n}{2}\alpha \gamma) {\ }Y_{00,i}{\ }{\ }{\ }{\ }(i=1,2,3) $$ which can be written in the form, $$ \frac{d^2x^i}{dt^2} = - \pder{\Phi_s}{x^i} {\ }{\ }{\ }{\ }(i=1,2,3) {\ }{\ }, \eqno{(5.7)} $$ where $$ \Phi_s = \frac{c^2}{2}{\ }\epsilon{\ }(1 -\frac{n}{2}\alpha \gamma) {\ }Y_{00}. \eqno{(5.8)} $$ Equantion (5.7) has the same form as Newton's equation of motion of a particle in a gravitational field having the potential $\Phi_s$ given by (5.8), which differs from the classical Newtonian potential. In the case of macroscopic particles $(n=0)$, we get from (5.8): $$ \Phi_s = \frac{c^2}{2}{\ }\epsilon{\ }Y_{00} = \Phi_N \eqno{(5.9)} $$ where $\Phi_N$ is the Newtonian gravitational potential obtained from a similar treatment of the geodesic equation. Thus (5.8) can be written in the form $$ \Phi_s = (1 - {\frac{n}{2}} \alpha \gamma) \Phi_N. \eqno{(5.10)} $$ This last expression shows that the gravitational potential felt by the spinning particle is less than that felt by a spinless particle or a macroscopic test particle. In other words, the Newtonian potentional is reduced for spinning particles, by a factor $(1 - {\frac{n}{2}} \alpha \gamma)$. \section{Summmary and Discussion} The marriage between the two philisophical ideas of the present century, quantization and geometrization, has never been successful so far. It is well known that quantization is successful in the domain of microphysics, while geometrization is successful in the domain of macrophysics, astrophysics and cosmology. For example, the dynamics of microscopic particles is well described within the framework of quantization, while the dynamics of macrophysical systems is described successfully in the framework of geometrization. The question now is: what is the best description of the dynamics of microscpic particle in a background gravitational field? Several authors have tried to answer this question, as mentioned in section 1. The problem is that their trails neither represent quantization of geomerty, nor represent geometrization of quantum mechanics. A solution of this problem may be, either starting by a certain geometry and using an appropriate procedure to see whether we can discover quantum features in this geometry; or starting in the quantum domain and using an appropriate procedure to look for geometric features. We believe that any appropriate geometry describing nature should contain paths that are quantized naturally, i.e.without using any quantization scheme, but how to discover such paths? The trial given in the present work represents a step in this direction. Although this trial is still far beyond quantization of geometry, it gives a strong evidence that paths in the AP-geometry are, in some sense, quantized. We believe that this result is valid for any non-symmetric geometry (geometry with torsion), but this statement needs confirmation. Furthermore one can consider the present work as a step twards unification of the dynamics of microscopic and macroscopic particles, since the new path equation (3.5) could be applied for, macroscopic or microscopic, massive or massless, and for spinning or spinless particles. AP-spaces are defined as spaces whose structure, in four dimensions, is defined completely by a tetrad vector field subject to the condition (2.3) (cf.Robertson (1932)). There is no complete agreement between authors on whether the AP-spaces are, in general, curved or flat. Because of (2.15) many authors believe that these spaces are flat (cf.Hayashi and Shirafuji (1979)). Others believe that these spaces are curved (cf.Mikhail and Wanas (1977)). The cause is that for a non-symmetric geometry (with a non-symmetric affine connexion) the curvature is not uniquely defined. However, all curvature tensors, in this space, are defined in terms of the torsion tensor, and consequently the vanishing of the torsion will leed to a flat space (Wanas and Melek(1995)). It is clear that the present work solves this controversal problem . The geometric structure establised in the present work has the following features :\\ (1) The structure of the space is defines completely (in 4-dimensions) by using a tetrad subject to the condition (2.3) and thus, by definition, we are still using an AP-geometry. So all what is done, from the gemetric point of view, is that some hidden structures in this geometry are illuminated. We are goining to call the structure developed in section 2 the Parametric Absolute Parallelism (PAP)-Space. \\ (2) It is an affinely connected space endowed by a general non-symmetric connexion (2.5). Thus the space possesses sufficient structure to carry out the operations of tensor analysis. \\ (3) At any point of the PAP- space we can define a metric tensor (2.2) which can be used to carry out the operations of raising and lowering indicies.\\ (4) The space is certainly curved since the curvature tensors (2.9) and (2.11) , corresponding to the non-symmetric connexion (2.5), are all non-vanishing tensors. It is to be considered that other curvature tensors could be defined in the PAP-space by generalizing the AP-derivatives. All these tensors are non vanishing and reduce to Riemann-Christoffel curvature tensor, of the Riemannian geometry, as b= 0. \\ (5) Paths in this geometry, corresponding to the connexion (2.5), are characterized by equation (3.5), which is considered as a generalization of the geodesic equation of Riemannian geometry. \\ (6) The new structure covers both the Riemannian geometry (a = 1, b = 0) and the conventional AP-geometry (a = 0, b = 1). Thus it is more general than both geometries. The new structure can be used to solve the problem (raised by Wanas and Melek(1995)) of constructing field theories, in non-symmetric geometries, that may be reduced to GR for $b=0$ without any need for vanishing torsion. The torsion of the space-time is assumed to be generated by rotating sources. Many authors believe in this statement (cf. Ross (1989)) . We beleive that torsion and metric tensors are both generic features of the gravitational field whether or not its source is rotating. Calculations show that, a sphecially symmetric solution for a version of general relativity written in AP-space, torsion has non-vanishing components while the source of the field is non-rotating (cf. Wanas (1990)). It is widely accepted that scalar particles (or macroscopic objects) cannot feel the torsion. As stated by some authors, scalar test particles detect the metric of the space while rotating test particles detect the torsion (cf. Nitch and Hehl(1980)). This is similar to the situation that neutral particles cannot feel the existence of an electromagnetic field. The question now is what is the intrinsic property of the test particle that interacts with the background field ? In case if background electromagnetic field the property is the electric charge. Similarly if the background field is gravitational then the intrinsic property is the mass-energy of the particle which interacts with the metric and/or the spin of the particle which interacts with the torsion. It has been shown (Wanas et al.(1995)b) that the results of applying the new path equation (3.5), in the solar system, are consistent with the observational bases of GR. The author would like to thank Dr.M.Melek for many discussions, and Mr.M.E. Kahil for checking relevant calculations. \section{References} Bazanaski, S. L. (1977) Ann. Inst. H. Poincar\'{e},A {\bf{27}}, 145. ~~~~~~~~~~~~~~~~~~(1989) J. Math. Phys. {\bf{30}}, 1018. \\ DeWitt, B. S. (1957) Rev. Modern Phys. {\bf{29}}, 377. \\ Galvao, C. and Teitelboim, C. (1980) J. Math. Phys. {\bf{21}}, 1863. \\ Hayashyi, K. and Shirafuji, T. (1979) Phys. Rev. D {\bf {19}}, 3524. \\ Hehl, F. W. (1971) Phys. Lett. {\bf{36}}A, 225. \\ Levi-Civita, T. (1950) "{\it {A Simplified Presentation of Einstein's United Field Equations}}", ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~ translated pamphlet, Blackie. \\ Longo, M. J. (1988) Phys. Rev. Lett. {\bf{60}}, 173. \\ Melek. M. (1988) Acta Physica Solvaca {\bf{38}}, 146. \\ Mikhail, F. I. (1952) Ph.D. Thesis, University of London.\\ Mikhail, F. I. and Wanas, M. I. (1977) Proc. Roy. soc. Lond. A {\bf{356}}, 471.\\ M{\o}ller, C. (1978) Mat. Fys. Medd. Dan. Vid. Selek. {\bf{39}},1. \\ Nitch, J. and Hehl, F. W. (1980) Phys. Lett. {\bf{90}}B, 98. \\ Robertson, H. P. (1932) Ann. Math. Princeton {\bf{33}}, 493. \\ Ross, D. K. (1989) Int. J. Theort. Phys. {\bf{28}}, 1333. \\ Wanas, M. I. (1990) Astron. Nachr. {\bf{ 311}}, 253. \\ Wanas M. I. and Melek M. (1995) Astrophys. Space Sci, {\bf{228}}, 277. \\ Wanas, M. I., Melek, M. and Kahil, M. E. (1995)a Astrophys. Space Sci,{\bf{228}}, 273. ~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~(1995)b to be submitted for publication. \end{document}
\section{Introduction} With the discovery of an apparent separation between the classical and quantum classifications of computational complexity \cite{shor}, and of fault-tolerant schemes for quantum computation \cite{shor_ft}, quantum information theory has earned a lasting and prominent place at the foundations of computer science. But at present this discipline seems rather isolated from most of the rest of physics. Will this change in the future? How might it change? One view is that thinking about information theory will lead us to a deeper understanding of the foundations of quantum mechanics. This vision has been vividly expressed by John Wheeler \cite{wheeler}; Bill Wootters \cite{wootters} and Chris Fuchs \cite{fuchs} have been among its particularly eloquent spokespersons. But I am not convinced in my heart that we are supposed to understand the foundations of quantum mechanics much better than we currently do. So I prefer to look in a different direction to anticipate where quantum information may have an impact on physics. What I tend to find most exciting in science are ideas that can build bridges across the traditional boundaries between disciplines. Perhaps that is why I find quantum computation appealing --- it has established an unprecedentedly deep link between the foundations of computer science and the foundations of physics. Truly great ideas in science tend to have broad consequences that can't be anticipated easily. Now the quantum information community is sitting atop two ideas with potential for greatness: quantum computation and quantum error correction. I'd like to suggest two directions in which quantum information theory might evolve in the future that could lead to broad and exciting consequences for other subfields of physics. These are: \begin{description} \item{1.} {\bf Precision measurement.} \newline Our deepening understanding of quantum information may lead to new strategies for pushing back the boundaries of quantum-limited measurements. Quantum entanglement, quantum error correction, and quantum information processing might all be exploited to improve the information-gathering capability of physics experiments. \item{2.} {\bf Many-body quantum entanglement.} \newline The most challenging and interesting problems in quantum dynamics involve understanding the behavior of strongly-coupled many-body systems --- systems with many degrees of freedom that undergo large quantum fluctuations. Better ways of characterizing and classifying the features of many-particle entanglement may lead to new and more effective methods for understanding the dynamical behavior of complex quantum systems. \end{description} \section{Quantum information theory and precision measurement} The connections between quantum information and precision measurement are explored in a separate article \cite{childs}, which I will only summarize here. My own interest in the quantum limitations on precision measurement has been spurred in part by Caltech's heavy involvement in the LIGO project, the Laser Interferometer Gravitational-Wave Observatory \cite{ligo}. LIGO is scheduled to begin collecting data in 2002, and a major upgrade is planned for two years later, which will boost the optical power in the interferometer and improve the sensitivity. In its most sensitive frequency band, the LIGO II observatory will actually be operating at the standard quantum limit (SQL) for detection of a weak classical force by monitoring a free mass. (In this case, the SQL corresponds to a force that nudges an 11 kg mass by about $10^{-17}$ cm at a frequency of 100 Hz.) Then within another 4 years (by 2008), another upgrade is expected, which will boost the sensitivity in the most critical frequency band beyond the SQL. Even an improvement by a factor of two can have a very significant payoff, for a factor of two in sensitivity means a factor of 8 in event rate. But the design of the LIGO III detection system is still largely undecided --- clever innovations will be needed. So Big Science will meet quantum measurement in the first decade of the new century, and ideas from quantum information theory may steer the subsequent developments in detection of gravitational waves and other weak forces. I learned the right way to think about the quantum limits on measurement sensitivity from Hideo Mabuchi \cite{mabuchi} --- in a quantum measurement, a classical signal is conveyed over a quantum channel.\footnote{Of course, connections between quantum information theory and precision measurement have been recognized by many authors. Especially relevant is the work by Wootters \cite{wootters}, by Braunstein\cite{braun}, and by Braunstein and Caves \cite{braunstein} on state distinguishability and parameter estimation, and by Braginsky and others \cite{braginsky} on quantum nondemolition measurement.} Nature sends us a message, a weak classical force, that can be regarded as a classical parameter appearing in the Hamiltonian of the apparatus (or more properly, if there is noise, a master equation). The apparatus undergoes a quantum operation $\$(a)$, and we are to extract as much information as we can about the parameter(s) $a$ by choosing an initial preparation of the apparatus, and a POVM to read it out. Quantum information theory should be able to provide a theory of the {\sl distinguishability of superoperators}, a measure of how much information we can extract that distinguishes one superoperator from another, given some specified resources that are available for the purpose. This distinguishability measure would characterize the inviolable limits on measurement precision that can be achieved with fixed resources. I don't know exactly what shape this nascent theory of the distinguishability of superoperators should take, but there are already some highly suggestive hints that progress in quantum information processing can promote the development of new strategies for performing high-precision measurements. \subsection{Superdense coding: improved distinguishability through entanglement} A watchword of quantum information theory is: ``Entanglement is a Useful Resource.'' It should not be a surprise if entanglement can extend the capabilities of the laboratory physicist. For example, the phenomenon of superdense coding illustrates that shared entanglement can enhance classical communication between two parties \cite{wiesner}. The same strategy can sometimes be used to exploit entanglement to improve the distinguishability among Hamiltonians (an idea suggested by Chris Fuchs \cite{fuchs_sd}). Suppose I wish to observe the precession of spin-$1/2$ objects to determine the value of an unknown magnetic field. If two spins are available, one way to estimate the value of the unknown field is to allow both spins to precess in the field independently, and then measure them separately. An alternative method is to prepare an entangled Bell pair, expose one of the two spins to the magnetic field while the other is carefully shielded from the field, and finally carry out a collective Bell measurement on the pair. It turns out that in many cases (for example when we have no {\sl a priori} knowledge about the field direction), the entangled strategy extracts more information about the unknown field than the strategy in which uncorrelated spins are measured one at a time \cite{childs}. This separation still holds even if we allow the unentangled strategy to be {\sl adaptive}; that is, even if the outcome of the measurement of the first spin is permitted to influence the choice of the measurement that is performed on the second spin. \subsection{Grover's database search: improved distinguishability through driving} An important paradigm emerging from the recent studies of quantum algorithms is Grover's method for rapidly searching an unsorted database \cite{grover}. Farhi and Gutmann \cite{farhi} observed that Grover's algorithm may be interpreted as a method for improving the distinguishability of a set of Hamiltonians by adding a controlled driving term. In the formulation they suggested, the Hamiltonian acting in an $N$-dimensional Hilbert space is known to be one of the operators \begin{equation} H_x= E|x\rangle\langle x|~, \end{equation} where $\{|x\rangle, x=0,1,\dots,N-1\}$ is an orthonormal basis. We may gain information about the value of $x$ by preparing states, allowing them to evolve under $H_x$ for a while, and then measuring suitable observables. But determining the value of $x$ by this strategy requires a total time of order $N$. A more effective strategy is to modify the Hamiltonian by adding a controlled driving term \begin{equation} H_D= E|s\rangle\langle s| ~, \end{equation} where $|s\rangle=N^{-1/2}\sum_{y=0}^{N-1}|y\rangle$, so that the full Hamiltonian becomes $H'_x=H_x+H_D$. If the initial state $|s\rangle$ is prepared and allowed to evolve under $H'_x$ for a time $T=\pi\sqrt{N}/2E$, then an orthogonal measurement in the $\{|x\rangle\}$ basis will reveal the true identity of the Hamiltonian. The time required is of order $\sqrt{N}$; this is Grover's quadratic speed-up. In this Grover-Farhi-Gutmann problem, there is a sense in which an optimal measurement procedure is known: Just as the Grover iteration allows one to identify a marked state with a minimum number of queries to the oracle \cite{bbbv}, the Grover perturbation allows us to identify the actual Hamiltonian in the minimal elapsed time (asymptotically for large $N$). Grover's algorithm presumes the existence of a quantum oracle that can reply to coherent queries. In an algorithmic setting, the oracle may be regarded as a quantum circuit that can be executed repeatedly. In experimental physics, the quantum oracle is Nature, whose secrets we are eager to expose. The experimenter is challenged to find the most effective (and practical!) way to query Nature and learn her Truths. \subsection{Semiclassical quantum Fourier transform as adaptive phase measurement} Shor's quantum factoring algorithm \cite{shor}, which apparently achieves an exponential speed-up relative to classical algorithms, is based on the efficient quantum Fourier transform (QFT). Fourier analysis is a versatile tool in the laboratory, so we might expect the fast QFT to have important applications to physics. One example could be the high-precision measurement of a frequency, like the energy splitting between the ground state and an excited state of an atom \cite{childs}. As Cleve {\it et al.} \cite{cleve} have emphasized, the QFT can be viewed as a procedure for estimating an unknown phase. With a quantum computer, we could execute the quantum Fourier transform on $n$ two-level atoms, and then read out a result by measuring the internal state of each atom. If losses are negligible, the measurement outcomes provide an estimate of the frequency to an accuracy of order $2^{-n}$. This procedure makes optimal use of an essential resource (the number of atoms measured), in that about one bit of information about the frequency is acquired in each binary measurement. In fact, the complexity of the quantum information processing needed to execute this protocol is modest. In its ``semiclassical'' implementation proposed by Griffiths and Niu \cite{griffiths}, the QFT is an {\sl adaptive} procedure for phase estimation that makes use of the information collected in previous measurements to extract the best possible information from subsequent measurements. Less significant bits of the phase are measured first, and the measurement results are used to determine what single-qubit phase rotations should be applied to other qubits to extract the more significant bits more reliably. In conventional Ramsey spectroscopy, these single-qubit transformations are applied simply by prescribing the proper time interval between the Ramsey pulses. \bigskip These and other related examples give strong hints that ideas emerging from the theory of quantum information and computation are destined to profoundly influence the experimental physics techniques of the future. \section{\bf Many-body entanglement and strongly-coupled quantum physics} \subsection{Some signposts in Hilbert space} The most challenging and interesting problems in quantum mechanics concern many-body systems with strong quantum fluctuations. An important goal is to understand the dynamics of such systems, but it is not easy. Indeed, it is largely because strongly-coupled quantum dynamics is so difficult to understand that we want so badly to build a quantum computer \cite{feynman}! I expect that, short of building a full-blown quantum simulator, there are many possible theoretical advances that potentially could enhance our understanding of strongly-coupled systems, including advances that could emerge from the theory of quantum information. A central task of quantum information theory has been to characterize and quantify the entanglement of multipartite systems. Up until now, most attention has focused on systems divided into a small number of parts (like two\footnote{See \cite{ibm}, for example.}), but also of great importance are the properties of $n$-body entanglement in the limit of large $n$. Studies of these properties may give us some guidance concerning what quantum simulation problems are genuinely computationally difficult, and may suggest to experimenters what kinds of systems are most likely to exhibit qualitatively new phenomena. Hilbert space is a big place \cite{caves_fuchs}, and so far we have become familiar with only a tiny part of it. In its unexplored vastness, there is sure to be exciting new physics to discover. But much of Hilbert space is bound to be very boring indeed, so we will need some clear signposts to show the way to the exotic new phenomena. It is truism (but still profoundly true!) that More is Different \cite{anderson}. So many of the collective phenomena exhibited by many-body systems (crystals, phase transitions, superconductivity, fractional quantum Hall effect, $\dots$) would be exceedingly hard to predict from first principles. That's good news for experimenters --- marvelous things could happen in many-body systems that we have been unable to anticipate. But it is easier to find something new when theory can provide some guidance. \subsection{Quantum error-correcting codes} A prototype for many-body entanglement has been developed in the past few years: the quantum error-correcting codes \cite{qec}. For example, a (nondegenerate) code that can correct any $t$ errors in a block of $n$ qubits has the property that no information resides in any set of $2t$ qubits chosen from the block -- the density matrix of the $2t$ qubits is completely random. Information {\sl can} be encoded in the block, but the encoded information has a {\sl global} character; there is no way to access any information at all by looking at only a few qubits at a time. For example, associated with the familiar five-qubit code \cite{ibm,five_qubit} that can protect a single encoded qubit from an error afflicting any of the five qubits in the code block, there is a maximally entangled six-qubit pure state. This state has the property that if we trace over any three of the qubits, the density matrix of the remaining three is a multiple of the identity. It has been recognized only rather recently how unusual this state is \cite{gf4}: there exist no $2n$-qubit states with $n$ larger than three such that tracing over half of the qubits leaves the other half in a completely random state.\footnote{But there {\sl are} such maximally entangled states with more than six parts if each part is a higher-dimensional system rather than a qubit \cite{gottesman}.} Asymptotically, we don't know precisely ``how entangled'' an $n$-qubit state can be, but there are useful upper and lower bounds. For large $n$, the number $s$ of qubits such that the density matrix for any $s$ of the $n$ is random, {\sl must} satisfy $s/n < 1/3$ \cite{rains}. On the other hand, states with this property are known to exist for $s/n \le .1893\dots$ \cite{att}. Somewhere between $1/3$ and $.1893$, there is a critical value that has not yet been pinned down. These upper and lower bounds are instructive examples of interesting results regarding multi-body entanglement that have emerged from the study of quantum error-correcting codes. \subsection{Classes of entangled states} This kind of global encoding of information is actually found in some systems that can be realized in the laboratory, such as systems that exhibit the fractional quantum Hall effect \cite{stone}, or certain kinds of frustrated antiferromagnets. These systems have in common that the microscopic degrees of freedom are locally ``frustrated'' -- that is, they are unable to find a configuration that satisfactorily minimizes the local energy density. In response, the system seeks an unusual collective state that relieves the frustration, a state such that the microscopic degrees of freedom are profoundly entangled. Condensed matter physicists have found useful ways to characterize the global properties of the entanglement that results. For example, in the case of a two-dimensional system, we may consider how the ground state degeneracy of the system behaves on a topologically nontrivial surface in the thermodynamic limit. As Wen \cite{wen} emphasized, in fractional quantum Hall systems the degeneracy increases with the genus (number of handles) of the surface as \begin{equation} {\rm ground ~state ~degeneracy}\sim (A)^{\rm genus} ~. \end{equation} This dependence arises from the ``winding'' of entanglement around the handles of the surface, and the value of $A$ distinguishes qualitatively different types of entangled states that must be separated from one another by phase boundaries. Just such a topological degeneracy is exploited in the ingenious quantum error-correcting codes constructed by Alexei Kitaev \cite{kitaev}. A closely related observation is that in a two-dimensional system with a boundary, there can be excitations confined to the boundary, and the properties of these edge excitations reflect the nature of the entanglement in the bulk system \cite{wen_edge}. I am hopeful that quantum information theory may lead to other as yet unknown ways to characterize the entangled many-body ground states of condensed matter systems, which may suggest new types of collective phenomena. We should also advance our understanding of how the profoundly entangled systems that Nature already provides might be exploited for stable storage of quantum information. \subsection{Information and renormalization group flow} The renormalization group (RG), one of the most profound ideas in science, is another topic that might be profoundly elucidated by an information-theoretic approach. Especially in the hands of Ken Wilson \cite{epsilon}, the RG spawned one of the central unifying insights of modern physics, that of {\sl universality} --- physics at long distances can be quite insensitive to the details of physics at much shorter distances. Indeed, for the purpose of describing the long-distance physics, all of the short-distance physics can be absorbed into the values of the parameters of an {\sl effective field theory}, where the number of parameters needed is modest if we are content with predictions to some specified accuracy. So it is that physics is possible at all. Fortunately, it is not necessary to grasp all the subtleties of quantum gravity at the Planck scale to understand (say) the spectrum of the hydrogen atom in great detail! The renormalization group describes how a quantum field theory ``flows'' as we ``integrate out'' short distance physics, obtaining a new theory with a smaller value of the ultraviolet momentum cutoff $\Lambda$. ``Universal'' features are associated with the ``fixed'' points in the space of theories where the flow is stationary. In the neighborhood of each fixed point are a finite number of independent directions in theory space along which the flow is repelled by the fixed point, the ``relevant'' directions of flow. Each fixed point provides a potential description of physics in the far infrared, with the number of free (``renormalized'') parameters in the description given by the number of relevant directions of flow away from the fixed point. Infrared theories with more parameters are less generic, in the sense that more ``bare'' parameters in the microscopic Hamiltonian of the system need to be carefully tuned in order for the flow to avoid all relevant directions and hence carry the theory to the vicinity of the fixed point. Typically, RG flow will carry a theory from the vicinity of a less generic fixed point toward the vicinity of a more generic fixed point. Now there is at least a heuristic sense in which information is lost as a theory flows along an RG trajectory --- the infrared theory ``forgets'' about its ultraviolet origins. One of the most intriguing challenges at the interface of physics and information is to make this connection more concrete.\footnote{For an interesting recent attempt, see \cite{brody}.} Can we quantify how much information is discarded when a theory flows from the vicinity of one fixed point to the vicinity of another? The proposal that an effective theory forgets more and more about its microscopic origins under RG flow leads to a robust expectation. RG flow should be a {\sl gradient} flow: it always runs downhill (toward ``less information''), and never uphill (toward ``more information''). Indeed, this property {\sl does} hold for translationally invariant and relativistically invariant quantum field theories in one spatial dimension. Zamalodchikov's $c$-{\sl theorem} \cite{ctheorem} identifies a function $C$ of the parameters in the Hamiltonian that can extracted from the two-point correlation function of the conserved energy-momentum tensor, and shows that $C$ is non-increasing along an RG trajectory. At a fixed point, the quantity $C$ coincides with the central charge $c$ that characterizes the representation of the conformal algebra according to which the fields of the fixed-point theory transform. Last year, an extension of this result to higher even-dimensional spacetimes was reported \cite{c4d} (following a suggestion by Cardy \cite{cardy_c}). It seems natural that the Zamolodchikov $C$-function should have a sharp interpretation relating it to loss of information along the flow, but none such is known (at least to me). A more precise information-theoretic interpretation of RG flow might guide the way to more general formulations of the $c$-theorem, applicable for example to theories in odd-dimensional spacetimes and to theories with less symmetry. And it might enrich our understanding of the classification of fixed-point theories and the general structure of renormalization group flow. \subsection{Bulk-boundary interactions} If the information-theoretic foundations underlying the $c$-theorem continue to prove elusive, there is another related problem that might turn out to be more tractable. It is known that a one-dimensional system with a {\sl boundary} (like a semi-infinite antiferromagnetic spin chain) can sometimes exhibit an anomalous zero-temperature entropy. The entropy has a piece proportional to the length of the chain that vanishes as $T\to 0$, but there is also a length-independent contribution that is nonvanishing at zero temperature (discovered by Cardy \cite{cardy} and by Affleck and Ludwig \cite{affleck}). Ordinarily, we expect that zero-temperature entropy has an interpretation in terms of ground-state degeneracy, but in these systems (which have no mass gap, so that the ground-state degeneracy becomes a subtle concept in the thermodynamic limit), $g=e^{S(T=0)}$ is not an integer; hence the interpretation of the entropy is obscure. A fascinating feature is that the ``ground-state degeneracy'' $g$ is a {\sl universal} property --- in the vicinity of an RG fixed point, its value is insensitive to the ultraviolet details (the microscopic interactions among the spins in the chain). Furthermore, there is evidence for a $g$-{\sl theorem}; $g$ has a smaller value at more generic fixed points and a larger value at less generic fixed points \cite{affleck}. The $g$-theorem, like the $c$-theorem, invites an interpretation in terms of loss of information along an RG trajectory. But I am hopeful that the information-theoretic origin of the $g$-theorem may turn out to be easier to understand. Upon hearing of entropy at zero temperature, a quantum information theorist's ears prick up -- it sounds like entanglement. It is tempting to interpret the entropy as arising from entanglement of degrees of freedom isolated at the boundary of the chain with degrees of freedom that reside in the bulk. So far, I have been unable to find a precise interpretation of this sort, but I still suspect that it could be possible. \subsection{Holographic universe} While on the subject of bulk-boundary interactions, I should mention the most grandiose such interaction of all. A new view of the quantum mechanics of spacetime is emerging from recent work in string theory, according to which the quantum information encoded in a spatial volume can be read completely on the surface that bounds the volume (``the holographic principle'') \cite{thooft}. This too has a whiff of entanglement -- for we have seen that in a profoundly entangled state the amount of information stored locally in the microscopic degrees of freedom can be far less than we would naively expect. (Think of a quantum error-correcting code, in which the encoded information may occupy a small ``global'' subspace of a much larger Hilbert space.) The holographic viewpoint is particularly powerful in the case of the quantum behavior of a black hole. The information that disappears behind the event horizon can be completely encoded on the horizon, and so can be transferred to the outgoing Hawking radiation \cite{hawking} that is emitted as the black hole evaporates. This way, the evaporation process need not destroy any quantum information. As the evidence supporting the holographic principle mounts \cite{maldacena}, an unsettling question becomes more deeply puzzling: If quantum information can be encoded completely on the boundary, why does physics seem to be local? It's strange that I imagine that I can reach out and embrace you, when we are both just shadows projected on the wall. Perhaps as the tools for analyzing many-body entanglement grow more powerful, we can begin to grasp the origin of the persistent illusion that physics is founded on the locality of spacetime.\footnote{A different possible connection between quantum error-correcting codes and the black-hole information puzzle was suggested in \cite{preskill_ft}.} \section{Conclusions} In the future, I expect quantum information to solidify its central position at the foundations of computer science, and also to erect bridges that connect with precision measurement, condensed matter physics, quantum field theory, quantum gravity, and other fields that we can only guess at today. I have identified two general areas in which I feel such connections may prove to be particularly enlightening. Progress in quantum information processing may guide the development of new ideas for improving the information-gathering capabilities of physics experiments. And a richer classification of the phases exhibited by highly entangled many-body systems may deepen our appreciation of the wealth of phenomena that can be realized by strongly-coupled quantum systems. \acknowledgments My work on the applications of quantum information theory to quantum-limited measurements has been in collaboration with Andrew Childs and Joe Renes \cite{childs}. I'm very grateful to Hideo Mabuchi for stimulating my interest in that subject, and to Dave Beckman and Chris Fuchs for their helpful suggestions. I have also benefitted from discussions about precision measurement with Constantin Brif, Jon Dowling, Steven van Enk, Jeff Kimble, Alesha Kitaev, and Kip Thorne. I thank Michael Nielsen for emphasizing the relevance of quantum information in quantum critical phenomena, Ian Affleck for enlightening correspondence about conformal field theory, Anton Kapustin for a discussion about Ref. \cite{c4d}, Dorje Brody for informing me about Ref. \cite{brody}, and Curt Callan for encouragement. Finally, I am indebted to Ike Chuang for challenging me to speculate about the future of quantum information theory. This work has been supported in part by the Department of Energy under Grant No. DE-FG03-92-ER40701, and by DARPA through the Quantum Information and Computation (QUIC) project administered by the Army Research Office under Grant No. DAAH04-96-1-0386.
\section{Introduction} It is one of the great dreams in the field of few-nucleon physics to find a quantitative correct and theoretically reasonable three nucleon force(3NF). In the past many 3NF models were developed\cite{Fujita,Tuscon,Coon81,Brazil,TEXAS,Ruhr,Urbana}. An especially prominent one is the meson-theoretical 3NF, for instance in the form of the Tucson-Melbourne model (TM)\cite{Tuscon,Coon81}. The reason for studying 3NFs is the existence of disagreements between the 3N data and the theoretical predictions with NN forces only. First of all the theoretical binding energy of $^3$H lacks about 500-800keV in relation to the experimental value of 8.48MeV using recent realistic potentials (e.g. CD-Bonn\cite{cdbonn}, AV18\cite{av18}, Nijmegen 93, Nijmegen I, II\cite{nijm}). These potentials describe all 2N observables to a degree of accuracy of $\chi ^2 / N_{data} \sim$1. In the low energy three nucleon continuum we have demonstrated\cite{Gloeckle96} that most of the observables agree well with the data using nucleon-nucleon (NN) forces only, however, there are exceptions. Some of them are well known as the ``$A_y$ puzzle''\cite{Koike,Witala3,Witala,Hueber,Kievsky}. In the high energy region the theoretical predictions differ visibly from the data if one only takes NN forces into account. The ``Sagara discrepancy''\cite{Sagara,Koike2,Witala2,Sakamoto,Nemoto} is an example of this. These problems definitely require a new Hamiltonian in the realm of the three nucleon system. Moreover, the $A_y$-puzzle requires not only a $2\pi$-exchange 3NF to become explained\cite{Gloeckle96,Kievsky} but other 3NF mechanisms as well. Beside these low-energy discrepancies in the 3N continuum there are also discrepancies at higher energies. This can be expected naively due to the shorter range nature of the 3NF in comparison to the NN force. Recently it became possible to explain discrepancies between experiment and predictions using NN forces only for the neutron-deuteron (nd) total cross section \cite{Total} and the nd elastic differential cross section \cite{Diff} with the $2\pi$-exchange TM 3NF. From the point of view of chiral perturbation, this special category of the TM 3NF should be modified\cite{Friar99}. Chiral perturbation theory has been successfully applied in the $\pi N$ system \cite{piNchiral,piNchiral2} and it is already playing an important role in its application to the NN system as well\cite{NNchiral,NNchiral2,NNchiral3} In \cite{Friar99} it recommended that the pion range - short range part of the $c$-term of the TM $2\pi$-exchange 3NF should be dropped, based on arguments from chiral perturbation theory. In doing this the pion range - short range part of the $c$-term remains, which is of the same type than the $a$-term. This leads to a redefinition of $a$. The such modified TM 3NF,called TM', has essentially the same effects on continuum\cite{Hueber2} than the original TM 3NF. Much remains to be investigated in the relation between NN and 3NF's. In the next section we present calculations for the triton binding energy based on variations of the values of the strength parameters in the TM 3NF, individually and combined. The summary and the outlook are give in Section 3. \section{Variations of the Tucson-Melbourne 3NF and their Triton Binding Energies} The TM force has the operator form: \begin{eqnarray} V_{TM}^{(3)} ={ 1 \over{ (2\pi)^6 }} { {g^2 _{\pi NN} }\over { 4 m^2 }} { { F^2_{\pi NN} (\vec q^2 ) F^2_{\pi NN}(\vec {q'}^2) \vec \sigma_ 1 \cdot \vec q \vec \sigma _2 \cdot \vec {q'} } \over { ( \vec q ^2 + m_\pi^2 ) ( \vec {q '} ^2 + m_\pi ^2 ) }} [ O^{\alpha \beta } \tau_\alpha \tau_\beta ], \end{eqnarray} \begin{eqnarray} O^{\alpha \beta } =\delta ^{\alpha \beta} [ a + b \vec q \cdot \vec {q'} + c (\vec q ^2 + \vec {q'} ^2 )] -d ( \tau _3 ^\gamma \epsilon ^{ \alpha \beta \gamma } \vec \sigma _3 \cdot \vec q \times \vec {q'} ). \label{operator} \end{eqnarray} where $m_\pi$, $m$, $g _{\pi NN}$ and $ F_{\pi NN} (\vec q^2 )$ are the pion mass, the nucleon mass, the $\pi NN$ coupling constant and the vertex function , respectively. The superscript (3) denotes that this expression is only one of three cyclically permuted parts of the total TM 3NF. There are four parameters ($a$, $b$, $c$ and $d$ ) which are chosen according to certain physical concepts\cite{Tuscon,Coon81}. For practical calculations one needs to introduce the vertex function which is normally chosen as \begin{eqnarray} F_{\pi NN}(\vec q ^2) = { { \Lambda^2 - m_\pi ^2 } \over { \Lambda ^2 + \vec q ^2 }} \end{eqnarray} The triton binding energy turns out to be strongly dependent on the cut-off parameter $\Lambda$. In a phenomenological approach it can be used as a fit parameter to adjust the triton binding energy and this separately for each NN potential\cite{Stadler,Nogga}. Using these cut-off parameters we calculated the polarization transfer parameter $K_y^{y'}$ in the three-body continuum. While the individual pure NN force predictions are different they essentially coincide if those individually adjusted 3NFs where included and that prediction agrees rather well with the data\cite{Hempen98}. This is one example out of several where scaling with the triton binding energy exists for 3N continuum observables. In those studies we kept the original TM parameters ($a$, $b$, $c$ and $d$) and only varied the form-factor cut-off parameter$\Lambda$. Now we want to go one step further and study phenomenologically the dependence of the triton binding energy on the individual terms in the TM 3NF operators connected to the $a$-, $b$-,$c$- and $d$-term. Like in \cite{Nogga} we solve the Faddeev equation rigorously including the 3NF. We choose CD-Bonn as the NN interaction. The original parameters\cite{Coon81} of the TM model are given in Table 2.1. \begin{table} \caption{ Parameters for the original 3NF. \label{table1}} \begin{tabular}{lccccccc} $a $ [$m_\pi^{-1}$] & $b$ [$m_\pi^{-3}$] & $c$ [$m_\pi^{-3}$] & $d$ [$m_\pi^{-3}$] & $m_\pi$ [MeV] & $m$ [MeV] &$g ^2 _{\pi NN}$ & $\Lambda $ [$m_\pi$] \\ \hline 1.13 & -2.58 & 1.00 & -0.753 & 139.6 &938.926 &179.7 & 4.856 \\ \end{tabular} \end{table} The cut-off parameter $\Lambda$ is not the original one but adjusted to reproduce the triton binding energy together with CD-Bonn. We multiply each one of the parameters $a$, $b$, $c$ and $d$ by a factor $X$ ($ 0 \le X \le 1.5$ ), one after the other: \begin{eqnarray} \pmatrix{ a \cr b \cr c\cr d\cr} \longrightarrow \cases{ \pmatrix{ a X\cr b \cr c\cr d\cr} & case a \cr \pmatrix{ a \cr b X \cr c\cr d\cr} & case b \cr \pmatrix{ a \cr b \cr c X\cr d\cr} & case c \cr \pmatrix{ a \cr b \cr c\cr d X \cr} & case d \cr } \end{eqnarray} and determine the 3N binding energy for these four cases. The results are shown in Fig.~1. \begin{figure}[hbt] \input{all.ps} \caption{The binding energy of the triton. The star ``$\star$'' is a new solution for the $c$-parameter. } \label{fig1} \end{figure} We see that the parameter $a$ contributes negligibly to the 3N bound state and its presence or absence is unimportant. This explains why the prediction for the triton binding energy for the TM and TM' 3NFs are close to each other\cite{Hueber2}. The $b$- and $d$-terms however are important. The binding energy increases monotonically with their strength. Interestingly, the behaviour of the $c$-term is such that there are two solutions which lead to the experimental value. We find 0.150 as the new solution (see ``$\star$'' in Fig.1). Now, with the exception of $a$ we let all parameters float. The parameter $a$ is kept at its original value 1.13. We search for the sets ($b$, $c$, $d$) which fulfil the condition to produce the experimental binding energy. Thus we have now three independent variables $X$: \begin{eqnarray} \pmatrix{ a \cr b \cr c\cr d\cr} \longrightarrow \pmatrix{ a \cr b X_b \cr c X_c \cr d X_d \cr} \end{eqnarray} \begin{figure}[hbt] \input{para1.ps} \caption{Contour plot of the parameters $b$, $c$, $d$ for constant binding energy. } \label{fig2} \end{figure} Fig. ~2 shows contour plots for different $X_c$ and $X_b$ while keeping $X_d$ at different fixed values. Each line in Fig.~2 corresponds to the same binding energy (8.48MeV). The black spot indicates the position for the original values of the parameters ($X_b$=$X_c$=$X_d$=1). The star is as in Fig. 1. We see that these two solutions for $c$ found above lie on the line in Fig.~2. The $b$- and $c$- values to the left (right) of a curve for a particular value of $d$ lead to over-binding (under-binding). A crucial rule\cite{Fujita} corresponding to the properties of the $\Delta$ particle excitation mechanism is that the ratio $ b/d $ is 4. The Urbana-Argonne\cite{Urbana} 3NF follows this rule, since except for a phenomenological short range term it includes only the $2\pi$-exchange with an intermediate $\Delta$ isobar, the Fujita-Miyazawa 3NF. The TM value for the ratio $b/d$ is 3.43, since in this model the $b$- and $d$-term include other processes on top of the $2\pi$-exchange with an intermediate $\Delta$. Also $b$ and $d$ are larger for the TM 3NF than for the pure $\Delta-2\pi$-exchange. The values for $b$ and $d$ of the Brazil and RuhrPot 3NF are close to those of the TM 3NF. The Texas 3NF, based on chiral perturbation theory, has even larger values for $b$ and $d$. This shows that it is not at all clear which values for the strength parameters within the 3NF one should choose. In Fig.~3 we show the curve for $b$=4$d$ and of course the additional requirement that the triton binding energy has the experimental value. As in Fig.~2 the underlying NN potential is CD-Bonn. Of course, in looking to Fig.~3 one should keep in mind that the choice for the $NN\pi$ form-factor (4) leads to a very strong dependence of the strength of the 3NF on the cut-off parameter $\Lambda$, as well. From Table II in \cite{Friar99} the locations of the $b$ and $c$ parameters for several 3NFs are indicated. Note however those 3NFs are not adjusted to the triton binding energy together with the CD-Bonn potential. \begin{figure}[hbt] \input{para2.ps} \caption{Parameters for various 3NFs.} \label{fig3} \end{figure} \section{Summary and outlook} Stimulated by \cite{Friar99} we studied the binding energy of $^3 $H as a function of the strength parameters ($a$, $b$, $c$ and $d$ in (\ref{operator}) ) in the TM force. We find that the $a$-term is not decisive when varying in the interval $ 0 \le a \le 2 $. The $b$- and $d$-terms, however are very important to obtain the experimental value 8.48MeV. Varying $c$ from 0 to 1.5 $\times c_{TM}$ we find that there are two solutions which belong to the same binding energy. The new solution is 15\% of the original value, namely, 0.15 [$m_\pi^{-3}$]. It supports phenomenologically the recommendation given in\cite{Friar99} ( based on arguments from chiral perturbation theory) that the short-range part of the $c$-term in the TM force should be dropped. If one assumes a purely phenomenological point of view for choosing the values for $a$ to $d$ in a 3NF of the form (\ref{operator}) Fig.~ 2 provides a complete overview for the possible values under the requirement that the triton binding energy is gained together with the CD-Bonn potential. Clearly corresponding pictures could be gained for other NN potentials. Of course other 3NF mechanisms have to be explored, too. At least for the $A_y$-puzzle it is clear that the $2\pi$-exchange 3NF is not sufficient to explain this discrepancy between theory and data. A study of pion range - short range 3NF terms is underway\cite{Hueber2} where are predicted by chiral perturbation theory. Based on the chosen form (2) and the requirement to fit the triton binding energy those 3NFs can now be tested in the 3N continuum with high energy. At IUCF\cite{IUCF}, RIKEN\cite{Sakamoto,Sakai} and KVI measurements are underway for 3N observables between 100-300 MeV. These are cross sections and various spin observables. They will be analysed using the 3NFs fixed in Fig.~2. This might allow to find a preference for a certain region in that parameter space or will show that additional forms are needed. \smallskip {\bf Acknowlegements} This paper is dedicated to Prof. Dr. Walter Gl\"ockle on the occasion of his 60th birthday. This work is financially supported by the Deutsche Forschungsgemeinschaft under Project No. Gl 87/19-2, No. Hu 746/1-3 and No. Gl 87/27-1, and partially the auspices of the U.S.\ Department of Energy. The calculations have been performed on the CRAY T3E of the John von Neumann Institute for Computing, J\"ulich, Germany.
\section{Introduction} \setcounter{equation}{0} As approximation schemes for gauge dynamics, instanton calculus \cite{shif} and 't Hooft's $1/N$ expansion \cite{tof} do not seem to combine in a useful fashion. Since effects of a charge $k$ instanton sector are of $\ord (e^{-8\pi^2 k/g^2}) = \ord (e^{-N})$, it would seem that they are always irrelevant in the large-$N$ limit unless they control the {\it leading} contribution to some observable (for instance, because of supersymmetry non-renormalization arguments), or somehow the integral over instanton moduli space is ill-defined. Such non-commutativity of the large-$N$ limit and the instanton sum is assumed to be behind well-known instances of theta-angle dependence at perturbative order in the $1/N$ expansion, notably in the context of large-$N$ chiral dynamics \cite{venewitt}. On the other hand, it is known that some toy models \cite{toym} completely suppress instanton-like excitations once the large-$N$ limit has been taken. In other words, the `effective action' resulting after large-$N$ diagram summation does not support instantons any more. So, one may wonder whether the large-$N$ `master field' always loses its discrete topological structure. Recently, the AdS/CFT correspondence of \cite{Malda,gkpw} has provided a new set of non-trivial master fields for some gauge theories. In particular, the theta-angle dependence of $\CN=4$ Super Yang--Mills $({\rm SYM}_4)$ in four dimensions can be studied in the large-$N$ expansion via perturbative Type IIB string theory in ${\bf AdS}_5 \times {\bf S}^5$. It is saturated by instantons, which appear in the supergravity description as D-instantons \cite{dinst}. So, the radical view that an instanton gas is incompatible with the large-$N$ limit is not vindicated in this case. In this paper, we investigate these questions in a QCD cousin introduced by Witten \cite{wit2}, which admits a supergravity description of its master field, while removing the constraints of extended supersymmetry and conformal invariance. More precisely, we would like to learn under what conditions some kind of topological configurations (instantons) still give the leading semiclassical effects of $\ord (e^{-N})$, even after the planar diagrams have been summed over. We shall focus on the most elementary case of the dilute limit, \mbox{{\em i.e.~}} the single-instanton sector. One description of this theory is in terms of $N$ D4-branes wrapped on a circle ${\bf S}^1_\beta$ of length $\beta$, with thermal boundary conditions. At weak coupling, the low-energy theory on the D4-branes is a perturbative five-dimensional Super Yang--Mills theory $({\rm SYM}_{4+1})$, which reduces to non-supersymmetric, $SU(N)$ Yang--Mills theory in four euclidean dimensions $({\rm YM}_4)$, at distance scales much larger than the inverse temperature $T^{-1}= \beta$. Since five-dimensional gauge fields originate from massless open strings, their coupling scales as $g_5^2 \sim g_s\,\sqrt{\alpha'}$ with $g_s = {\rm exp}(\phi_{\infty})$ denoting the string coupling constant. Therefore, the four-dimensional coupling at the cut-off scale $T$ is given by $g^2 \sim g_s \,T\,\sqrt{\alpha'}$. The weak-coupling description of instantons in this set up is in terms of D0/D4-branes bound states. Due to the Wess--Zumino coupling between the type IIA Ramond--Ramond (RR) one-form and the gauge fields on the D4-branes world-volume, $ {\cal L}} \def\CH{{\cal H}} \def\CI{{\cal I}} \def\CU{{\cal U}_{\rm WZ} = C_{\rm D0} \wedge F\wedge F$, a D0-brane `inside' a D4-brane carries the instanton charge. The action of an euclidean world-line wrapped on a circle ${\bf S}_{\beta}^1$ is \begin{equation}\label{d0oldaction} S_{{\rm D}0} = M_{{\rm D}0} \cdot \beta = {\beta \over \sqrt{\alpha'} \,g_s}= {8\pi^2 \over g^2} \equiv {8\pi^2 N \over \lambda} .\end{equation} Incidentally, this relation also fixes the numerical conventions in the definition of the four-dimensional coupling $g$. We have also introduced the standard notation for the large-$N$ 't Hooft coupling $\lambda \equiv g^2 N$. The moduli of these instantons are encoded in the quantum mechanical zero-modes of the D0--D0 and D0--D4 strings. For a standard compactification, the D0-branes (\mbox{{\em i.e.~}} the `instanton particles' of the gauge theory) describe standard instantons of $\CN=4$ ${\rm SYM}_4$ (see \cite{usinst} for some generalizations). If the circle breaks supersymmetry, the instanton fermionic zero modes should be lifted accordingly to mass of $\ord(T)$, and one should get essentially a Yang--Mills instanton with no fermionic zero modes. Other one-loop effects would incorporate the perturbative running of the coupling constant in the standard way. The supergravity framework for ${\rm SYM}_{4+1}$ at finite temperature is given by the black D4-brane solution \cite{wit2, mali}. The full metric in the string frame is:\footnote{See, for example, \cite{krev} and references therein for a review of metrics relevant to this paper.} \begin{equation} \label{fulmet} ds^2 = H_4^{-\frac{1}{2}} (h\,d\tau^2 + d{\vec y}^{\,2}) + H_4^{\frac{1}{2}} \,\left( dr^2 / h + r^2 \,d\Omega_{4}^2 \right) \end{equation} where \begin{equation} \label{prof} H_4= 1+ (r_{Q4}/r)^3 \,,\;\;\;\;\;\;\; h=1-(r_0 /r)^3. \end{equation} There are two length scales associated with this metric: the Schwarzschild radius, $r_0$, related to the Hawking temperature $T$ by $ T^{-1} = \beta = (4\pi/3)\, r_0 \,[H_4 (r_0)]^{1/2}$, and the charge radius $r_{Q4}$, given by \begin{equation} \label{crad} r_{Q4}^3 = - {\mbox{$\half$}} r_0^3 + \sqrt{{\mbox{$1\over 4$}} r_0^6 + \alpha'^{\,3}\,(\pi\, g_s\,N)^2}, \end{equation} In the Maldacena or gauge-theory limit, one scales $\alpha' \rightarrow 0$ with $r/\alpha' =u $ and $r_0 /\alpha' = u_0$ fixed. The new coordinate $u$ has dimensions of energy and the scaling properties of the Higgs expectation value. In this limit, only the combination \begin{equation} \label{nea} \alpha'^{\,2} \,H_4 \rightarrow {\pi \,g_s \,N\,\sqrt{\alpha'} \over u^3} = {\lambda \,\beta\over 8\pi\, u^3} \end{equation} is relevant. In the supergravity picture, the D4-branes have disappeared in favour of the `throat geometry' ${\bf X}_{\rm bb}$ (\ref{fulmet}), \mbox{{\em i.e.~}} we have no open strings and the description is fully gauge invariant. The black-brane manifold ${\bf X}_{\rm bb}$, with topology ${\bf R}^2 \times {\bf R}^{4} \times {\bf S}^4$, has a boundary at $u=\infty$ of topology ${\bf S}^1 \times {\bf R}^4$, which is interpreted as the ${\rm SYM}_{5}$ space-time (the $(\tau, {\vec y})$ space). The physical interpretation is that asymptotic boundary conditions for the supergravity fields at $u=\infty$ represent coupling constants of microscopic operators in the gauge theory \cite{gkpw}. The same boundary conditions are satisfied by the extremal D4-brane metric with thermal boundary conditions. This is the `vacuum' manifold, denoted ${\bf X}_{\rm vac}$, with topology ${\bf S}^1 \times {\bf R}^5 \times {\bf S}^4$, obtained from (\ref{fulmet}) by setting $u_0 =0$, with {\it fixed} $\beta$. However, one can show \cite{wit2,bkr} that ${\bf X}_{\rm vac}$ is suppressed by a relative factor of $\ord (e^{-N^2})$ with respect to ${\bf X}_{\rm bb}$, in the large-$N$ limit. In other words, the $\ord (N^2)$ actions satisfy \begin{equation} \label{balan} I({\bf X}_{\rm bb}) - I({\bf X}_{\rm vac}) = -K\,N^2 \,\lambda \, VT^4 <0 \end{equation} for any $T>0$, with $K$ a positive constant, \mbox{{\em i.e.~}} there is no Hawking-Page transition \cite{hp,wit2}. Unlike the case of $\CN =4$ ${\rm SYM}_4$, the dilaton is not constant in the supergravity description. It becomes strongly coupled at radial coordinates of order $u\sim \ord (N^{4/3} /\beta\,\lambda)$, where one has $e^{\phi} = g_s \,(H_4)^{-1/4} =\ord (1)$. Beyond this point, one should use a dual picture in terms of a wrapped M5-brane in M-theory, \mbox{{\em i.e.~}} a quotient of ${\bf AdS}_7 \times {\bf S}^4$. For the purposes of the discussions in this paper, we are studying the theory at fixed energy scales of $\ord (1)$ in the 't Hooft's large-$N$ limit, with fixed $\lambda=g^2 N$. Therefore, such non-perturbative thresholds effectively decouple in the regime of interest, and we shall formally extend the D4-brane manifold all the way up to $u=\infty$. {}From a physical point of view, $\alpha'$-corrections to the classical geometry pose a more serious limitation to the supergravity description. The curvature at the horizon scales as $( u_0 \,g_s\,N \,\sqrt{\alpha'})^{-1/2} \sim \lambda^{-1}$, in string units, so that the supergravity description is accurate only for large bare 't Hooft coupling $\lambda \gg 1$. On the other hand, the glueball mass gap \cite{oo} in this theory is of order $ M_{\rm glue} \sim \beta^{-1} =T$, while inspection of the Wilson loop expectation value gives a four-dimensional string tension \cite{stten} of order $\sigma \sim \lambda \,T^2$, \mbox{{\em i.e.~}} hierarchically larger in the supergravity regime. This lack of scaling indicates that the supergravity picture is far from the `continuum limit' of the ${\rm YM}_4$ theory, a suspicion already clear from the existence of non-QCD states of Kaluza--Klein origin at the same mass scale as the glueballs: $M_{\rm KK} \sim T \sim M_{\rm glue}$. \section{The Localized Instanton} \setcounter{equation}{0} The natural candidate for an instanton excitation in the large-$N$ supergravity picture is a D0-brane probe wrapped around the thermal circle. For the supersymmetric case, this is indeed the T-dual configuration to the D-instantons in ${\bf AdS}_5 \times {\bf S}^5$ discussed in \cite{dinst}. Wrapped D0-branes have the correct quantum numbers to be interpreted as Yang--Mills instantons in the effective four-dimensional theory. The topological charge is interpreted as the wrapping number on the thermal circle ${\bf S}^1_{\beta}$. In the large-$N$ limit, it is justified to take the D0-brane as a probe, neglecting its back-reaction on the supergravity fields, since the gravitational radius is sub-stringy: $(r_{\rm probe})^7 \sim \alpha'^{\,7/2}\,e^{\phi} \sim \ord (1/N)$, although for instanton numbers of $\ord (N)$ with identical moduli we may need a supergravity description for the instanton dynamics in terms of the D0-branes near-horizon geometry (\mbox{{\em i.e.~}} a T-dual of the limit in \cite{koo}, or the solution of section 3 below). We shall postpone these interesting complications by working in the single-instanton sector, and with instanton moduli of $\ord (1)$ in the 't Hooft large-$N$ limit. One important ingredient of the the instanton/D0-brane mapping is a physical interpretation in gauge-theory language of the wrapped D0 world-line's radial position. For this purpose, we use the generalized UV/IR connection as discussed in \cite{uvir}. According to this, a radial coordinate $u$ is associated to a length scale in the ${\rm SYM}_{p+1}$ gauge theory of order $ \ell\sim \sqrt{ g_{p+1}^2 \,N /u^{5-p}}$. Thus, in our case, the size parameter $\rho$ of the instanton satisfies: \begin{equation} \label{sizep} \rho^2 = {\beta\,\lambda\over u} .\end{equation} We will assume this relation as the definition of the instanton's size modulus. We will now discuss both manifolds with ${\bf S}^1 \times {\bf R}^4$ boundary conditions at $u=\infty$, in spite of the fact that eq.\ (\ref{balan}) ensures the dynamical dominance of ${\bf X}_{\rm bb}$. The reason for considering also the `vacuum' manifold is first that we find interesting differences between the manifolds, and that ${\bf X}_{\rm vac}$ is the only relevant manifold for supersymmetric compactification, with which we can make contact with the ${\bf AdS}_5 \times {\bf S}^5$ case. \subsection{Vacuum Manifold} The ${\bf S}^1$ factor on the boundary extends to the bulk of ${\bf X}_{\rm vac}$ becoming singular as $u\rightarrow 0$, since ${\rm Vol}({\bf S}^1_u) = \beta \sqrt{g_{\tau\tau}} = \beta\,(H_4)^{-1/4} \rightarrow 0$. However, the action of the instanton is constant, due to the dilaton dependence in the Dirac--Born--Infeld action: \begin{equation} \label{dbis} S_{\rm D0} = M_{\rm D0} \int_{0}^{\beta} d\tau\, (g_s\,e^{-\phi}) \,\sqrt{g_{\tau\tau}} = {8\pi^2 \over g^2} .\end{equation} Thus, the size $\rho$ is an exact modulus in the supergravity description on ${\bf X}_{\rm vac}$. On general grounds, the path integral of a D-particle in a curved background ${\bf X}$ contains an ultralocal term in the measure of the form $\CD X^{\mu} \,[{\rm det}(g_{\mu\nu})]^{1/2}$, to ensure invariance under target-space diffeomorphisms. In the description of instantons on the manifold ${\bf X}$, we concentrate on the zero-mode part which then leads to a single-instanton measure \begin{equation} \label{measurex} d\mu \,({\bf X}) = C_{N,\lambda} \;d\eta\;\int_{{\bf S}^1_{\beta} \times {\bf S}^4_{\Omega} } (\alpha')^{-5}\;d{\rm Vol}({\bf X}), \end{equation} where $d\eta$ is the measure over fermionic zero-modes (sixteen in the supersymmetric case), and $C_{N,\lambda}$ is a constant to be determined by the matching to the perturbative measure. We have produced a measure in the physical space-time and scale-parameter space by averaging over ${\bf S}^1_{\beta} \times {\bf S}^4_{\Omega}$. The result for ${\bf X}_{\rm D4} = {\bf X}_{\rm vac}$, using the UV/IR connection (\ref{sizep}) is: \begin{equation} \label{measud4} d\mu \,({\bf X}_{\rm D4}) = C_{N,\lambda} \; \lambda^5 \, (\rho \,T)^{-6} \; \rho^{-5}\,d\rho\,d{\vec y} \,d\eta. \end{equation} We see that the presence of the dimensionful scale $T$ explicitly violates the conformal invariance of the measure, which we must take as a concrete prediction of the supergravity approach. As such, it is valid at large $N$ and $\lambda$. The singularity of ${\bf X}_{\rm vac}$ as $u\rightarrow 0$ is not relevant. At $u\sim u_s = T \lambda^{1/3}$ the size of the world-line is of $\ord (1)$ in string units. So, for $u\ll u_s$ we must use the T-dual metric of $N$ D3-branes smeared over the dual circle of coordinate length ${\widetilde \beta} = 4\pi^2 \alpha' /\beta$: \begin{equation} \label{tdual} ds^2 ({\bf X}_{\widetilde {\rm D3}}) = H_4^{-\frac{1}{2}} \, d{\vec y}^{\,2} + H_4^{\frac{1}{2}}\, \left( d{\widetilde \tau}^2 + dr^2 + r^2 \,d\Omega_4^2 \right) ,\end{equation} with ${\widetilde \tau} \equiv {\widetilde \tau} + {\widetilde \beta}$. In the T-dual metric,\footnote{Notice that the UV/IR connection (\ref{sizep}) remains unchanged by T-duality, as the new metric only differs by $g_{\tau\tau} \rightarrow 1/g_{\tau\tau}$, with the $u,{\vec y}$ components of the metric unaffected.} the size of ${\bf S}^1_u$ grows with decreasing $u$. In fact, the metric (\ref{tdual}) is unstable if any small amount of energy is added. It collapses to the array solution of localized D3-branes \cite{greglaf}: $$ ds^2 ({\bf X}_{\rm D3}) = H_3^{-\frac{1}{2}} d{\vec y}^{\,2} + H_3^{\frac{1}{2}}\, \left(d{\widetilde\tau}^2 + dr^2 + r^2 \,d\Omega_4^2 \right), \;\;\;\;{\rm with} \;\;\; H_3 = 1+\sum_n {4\pi\,{\widetilde g}_s \,\alpha'^{\,2} \over |r^2 + (n{\widetilde\beta})^2 |^2}. $$ By the T-duality rules and our coupling conventions: $4\pi{\widetilde g}_s = 8\pi^2 \,g_s \sqrt{\alpha'}/\beta = g^2$. In the regime $r\gg {\widetilde \beta}$ we can approximate the discrete sum over images by a continuous integral, and we recover the smeared metric (\ref{tdual}) as an approximation. On the other hand, for $r\ll {\widetilde \beta}$ we can instead neglect the images and approximate the sum by the $n=0$ term. The result is of course the standard ${\bf AdS}_5 \times {\bf S}^5$ metric corresponding to D3-branes at strong coupling. Indeed, the UV/IR relation for D-instantons in D3-branes \cite{dinst}, $ \rho = \sqrt{\lambda} / u ,$ matches the five-dimensional one (\ref{sizep}) precisely at $u=u_{\rm loc} = 1/\beta$, which is equivalent to $r=r_{\rm loc} = {\widetilde \beta}/4\pi^2$. The instanton measure (\ref{measud4}) matches across these finite-size transitions to the corresponding measures for the new manifolds ${\bf X}_{\widetilde {\rm D3}}$ and ${\bf X}_{\rm D3}$, because the definition (\ref{measurex}) applies in general and the volume form matches across the transitions at $u=u_s$ and $u=u_{\rm loc}$. The resulting measures are (both up to $\ord (1)$ numerical factors): \begin{eqnarray} \label{meas} d\mu \,({\bf X}_{\widetilde {\rm D3}}) &=& C_{N,\lambda} \;\lambda^4 \;(\rho \,T)^{-3} \;\rho^{-5}\,d\rho\,d{\vec y}\,d\eta, \nonumber \\ d\mu \,({\bf X}_{\rm D3}) &=& C_{N,\lambda} \; \lambda^{5/2} \;\rho^{-5}\, d\rho\, d{\vec y}\, d\eta . \end{eqnarray} This last measure is conformally invariant, and coincides with that of refs. \cite{dinst} for D-instantons in ${\bf AdS}_5 \times {\bf S}^5$. Finally, as pointed out in the introduction, the validity of the supergravity picture is limited by the requirement that we can control the $\alpha'$-corrections. The curvature of the D4-brane metrics is of $\ord (1)$ in string units at the `correspondence line' $u_c \sim (\beta \,\lambda)^{-1}$ \cite{horpol}. For the D3-brane metrics, the condition is simply $\lambda \sim 1$. This implies that, for $\rho < \beta$, we have a correspondence line for instanton sizes $ \rho =\rho_c = \beta\,\lambda. $ For $\rho >\beta$ the correspondence line is independent of $\rho$ and lies at $\lambda \sim 1$. Below the correspondence line the system is better described in Yang--Mills perturbation theory, although we lose the analytic control over the $1/N$ expansion. The geometrical D-instanton measure in ${\bf AdS}_5 \times {\bf S}^5$ has been matched to the perturbative instanton measure in the $\CN=4$ ${\rm SYM}_4$ theory in great detail, including multi-instanton terms \cite{valya}. In particular, this allows us to fix the coupling-dependent constant as $C_{N,\lambda} = N^{-7/2} \, \lambda^{3/2}$. This is rather remarkable, since the geometrical measure holds at large $\lambda$, whereas the perturbative measure is derived in Yang--Mills perturbation theory, valid for $\lambda \ll 1$. This robustness of the instanton measure in this case might be due to the high degree of supersymmetry and/or conformal symmetry. For instance, the analogous matching between the D4-brane supergravity measure (\ref{measud4}) and the perturbative description of the `instanton particles' of ${\rm SYM}_5$, through the correspondence line $\rho = \rho_c = \beta \,\lambda$, fails by one power of $\lambda$. This means that the very precise matching of \cite{valya} for ${\bf AdS}_5$ D-instantons is probably a consequence of conformal invariance. This discussion may be summarized in Fig. 1, where the finite-size transitions, as well as the correspondence lines are depicted as a function of the 't Hooft coupling and the instanton size. \begin{figure} \hspace*{1.4in} \epsfxsize=3in \epsffile{ins1.eps} \caption{ \small Instanton phase diagram for the compactified D4 theory on a supersymmetric circle of size $\beta$. The dotted lines denote the correspondence curves separating the geometric descriptions at large 't Hooft coupling $\lambda =g^2 N$ from the perturbative SYM descriptions at small $\lambda$. Dashed lines represent transitions driven by the finite size of the compactification circle. } \end{figure} \subsection{Black-brane Manifold} Although the wrapping charge of D0-branes is well defined in ${\bf X}_{\rm vac}$, the thermal circle being non-contractible, this is not the case for ${\bf X}_{\rm bb}$, whose $(\tau, u)$ subspace has ${\bf R}^2$ topology. Therefore, the thermal circle at fixed radial coordinate ${\bf S}^1_u$, is contractible, being the boundary of a disc: ${\bf S}^1_u = \partial {\bf D}_u$, \mbox{{\em i.e.~}}\ we can `unwrap' the D0-brane instanton through the horizon. Thus, while exact instanton charges can be identified in the supersymmetric case, no quantized topological charge seems to survive in the non-supersymmetric case, due to the dynamical dominace of ${\bf X}_{\rm bb}$ in the large-$N$ limit (\ref{balan}). Still, we can talk of approximate or `constrained' instantons, provided the probe D0-brane world-line wraps far away from the horizon. In this case the un-wrapping costs a large action. In order to estimate the action as a function of $u$ (or the instanton size $\rho$), we calculate the Dirac--Born--Infeld action of the probe D0-brane: \begin{equation} \label{corac} S_{\rm D0} = M_{\rm D0} \int_{0}^{\beta} d\tau\, (g_s\,e^{-\phi}) \,\sqrt{g_{\tau\tau}} = {8\pi^2\over g^2} \sqrt{h} = {8\pi^2 \over g^2} \sqrt{1-(\rho/\beta)^6} ,\end{equation} where we have used the UV/IR relation (\ref{sizep}) in the last step. Thus, $\rho$ is not an exact modulus, as instantons tend to grow. For an instanton of the order of the glueball's Compton wave-length $\rho \sim \beta$, the action is comparable to the vacuum action, and the instanton has disappeared (un-wrapped). In the far ultraviolet, we can use the approximate instantons of very small size $\rho \ll \beta$, to measure a `running effective theta angle', by requiring that the approximate instanton is weighed by a phase ${\rm exp}(i\theta_{\rm eff})$, with $\theta_{\rm eff} (u=\infty)=\theta$, the bare theta angle of the four-dimensional ${\rm YM}_4$ theory. Following Witten \cite{thetaw}, a bare theta angle is associated to a RR two-form \begin{equation} \label{witprof} f_{\rm D0} = dC_{\rm D0} = {\overline \theta} \,{3\over \pi \zeta^7} \,d\zeta\wedge d\psi, \end{equation} where, in the notation of \cite{thetaw}, $\zeta^2 =u/ u_0$, and $\psi=2\pi\tau/\beta$. The bare theta angle, measured at $u=\infty$, is $\theta = {\overline \theta} \,({\rm mod} \;2\pi)$, due to the multiplicity of meta-stable vacua as described in \cite{thetaw}, \mbox{{\em i.e.~}} $f_{\rm D0} \propto {\overline\theta} = (\theta + 2\pi n)$ in the $n$-th vacuum (see also \cite{ozpa} for another geometric approach to this question). In what follows, we shall obviate this technicality by working in the $n=0$ vacuum, so that $\theta ={\overline \theta}$. The effective theta angle at throat radius $u$ is \begin{equation} \label{thetaf} \theta_{\rm eff} (u) = \oint_{{\bf S}^1_u} C_{\rm D0} = \int_{{\bf D}_u} f_{\rm D0} =\theta \left(1-6\int_{\zeta(u)}^{\infty} d\zeta \,\zeta^{-7}\right) = \theta \; h(u) = \theta \left(1-(\rho/\beta)^6 \right) .\end{equation} The `correspondence line' $u_c \sim (\beta \,\lambda)^{-1}$ \cite{horpol}, controlling $\alpha'$-corrections is also defined in ${\bf X}_{\rm bb}$. In terms of instanton sizes, for $\rho < \beta$, we have a correspondence line at $ \rho_c = \beta\,\lambda$. Since no instantons survive for $\rho > \beta$ in the supergravity picture, the finite-size effects related to T-duality in ${\bf S}^1_{\beta}$ and localization effects are absent for ${\bf X}_{\rm bb}$, \mbox{{\em i.e.~}} there is no phase of D-instantons in ${\bf AdS}_5 \times {\bf S}^5$. The situation can be summarized by Fig. 2. \begin{figure} \hspace*{1.4in} \epsfxsize=3in \epsffile{ins2p.eps} \caption{ \small Instanton phase diagram for the compactified D4 theory on a {\it thermal} circle of size $\beta$. We have continued the glueball mass scale curve $\rho \Lambda_{\rm QCD} \sim 1$ to weak coupling in a way tentatively consistent with asymptotic freedom. } \end{figure} \section{The Smeared Instanton Solution} \setcounter{equation}{0} In the previous section we have seen that probe D0-branes wrapping the thermal circle of a black D4-brane in the far ultraviolet are dual to (unstable) small-size instantons. Vice-versa, there exists a different supergravity solution dual to a field-theory configuration which can be interpreted as containing a condensate of {\it large\/} instantons. Indeed, the smeared, black D0/D4-brane solution is interpreted (as in ref.\ \cite{liutsey} for the supersymmetric T-dual case) to be dual to a Yang-Mills theory with a non-vanishing self-dual background. The self-duality of the background implies that it can be related to instantons, and the smeariness of the D0-branes can be interpreted very heuristically as the fact that the instantons are `smooth' and then `large'. In fact, in real time, the D0-branes are smeared on the D4-branes as soon as they `fall behind' the horizon, due to the no-hair property (this corresponds $u=u_0$ or, using (\ref{sizep}), to $\rho =\beta$.) This statement has only a heuristic value because, in the euclidean time configurations we are considering, space-time effectively ends at $u=u_0$. Still, the effects of the source D0-branes can be detected on the long-range fields such as the metric, dilaton, and RR fields. In this section, we pursue this view of the smeared D0-branes not as probes, as in the previous section, but as background data. The string-frame metric outside a system of $k$ D0-branes smeared over the volume of $N$ D4-branes differs from that in (\ref{fulmet}) by one more harmonic function $H_0$: \begin{equation} ds^2 = H_0^{-\frac{1}{2}} \,H_4^{-\frac{1}{2}} \,h\,d\tau^2 + H_0^{\frac{1}{2}}\, H_4^{ -\frac{1}{2}} \,d{\vec y}^{\,2} + H_0^{\frac{1}{2}} \,H_4^{\frac{1}{2}} \left(dr^2 / h+r^2 \,d\Omega_4^2\right) .\end{equation} In the gauge-theory limit, this function is given by \begin{equation} \label{acheo} H_0 (u) = 1+(u_{Q0}/u)^3= 1-{\mbox{$\half$}} (u_0 /u)^3 +\sqrt{{\mbox{${1\over 4}$}}(u_0 /u)^6 + (u_k /u)^6}. \end{equation} It depends on a new energy scale $u_k$, related to the number density of D0-branes $k/V\equiv \kappa$ by \begin{equation} \label{uk} u_k^3 = \kappa \,{2\pi^3\,\beta \,\lambda \over N} .\end{equation} The new scale is small $(u_k^3 = \ord (1/N))$ in the large-$N$ limit with fixed instanton charge density. In this paper, we are interested in the physics at energies of $\ord (1)$ in the large-$N$ limit, so that $u_k \ll u_0$ and $H_0$ may be approximated by~\footnote{At very low energies, $u_0 \ll u_k$, the smeared solution is ${\bf X}_6 \times {\bf T}^4$, with ${\bf X}_6$ conformal to ${\bf AdS}_2 \times {\bf S}^4$ in the sense of \cite{kostas}. It is presumably related to quantum mechanics in the large-$k$ instanton moduli space \cite{seimm}.} $ H_0 = 1 +(u_k^2 / u_0 \,u)^3 + \ord (1/N^4). $ The dilaton profile also receives $\kappa$-dependent corrections, $ g_s \,e^{-\phi} = (H_4 / H_0^3)^{1/4} $, as well as the Hawking temperature: \begin{equation} \label{hatt} T^{-1} = \beta= {4 \pi \over 3}\, r_0 \,\sqrt{H_0 (r_0) \, H_4 (r_0)} =\left({2\pi\lambda\beta\over 9\,u_0} \;H_0 (u_0)\right)^{1/2}.\end{equation} This yields an equation for $u_0$ that can be solved iteratively in powers of $(u_k /\lambda T)^6$. The relation between the smeared D0-brane number density $\kappa$ and the running theta angle is obtained from the supergravity solution for the RR two form: \begin{equation} \label{totp} f_{\rm D0}={c\,\kappa \over u^4} {1\over (H_0)^2} \,du \wedge d\tau ,\end{equation} with $c$ a known numerical constant. As before, a wrapped D0-brane probe can be used to measure an effective theta angle whose value at $u=\infty$ defines the bare theta angle. Plugging (\ref{totp}) into (\ref{thetaf}) we obtain: \begin{equation} \label{thotra} \theta_{\rm eff} (u) = \int_{{\bf D}_u} f_{\rm D0}= \theta \,{u^3 -u_0^3 \over u^3 + u_{Q0}^3}= \theta \,{h(u) \over H_0 (u)}, \qquad {\rm where}\qquad \theta = {\beta \,c\,\kappa \over 3 } {1\over u_0^3 + u_{Q0}^3} .\end{equation} The two-form solution found by Witten (\ref{witprof}) corresponds to the $u_0 \gg u_k$ regime of (\ref{totp}). This provides a relation between the number density $\kappa$ of smeared D0-branes and the bare theta angle, valid in the large-$N$ limit with fixed $\kappa$: \begin{equation} \label{match} \theta = {\beta \,c\,\kappa \over 3 \, u_0^3 \,H_0 (u_0)} =\kappa\cdot {9\,c\over \lambda^3 \,T^4} \cdot \left({3\over 2\pi}\right)^3 + \ord (1/N^2),\end{equation} where we have used $u_0 =2\pi \lambda T/9 + \ord (1/N^2)$, from (\ref{hatt}). In interpreting this relation, it is important to remember that we are working in the $n=0$ vacuum, out of the $\ord (N)$ metastable vacua mentioned in section 2.2, \mbox{{\em i.e.~}} the actual values of the parameters are such that the r.h.s of (\ref{match}) is smaller than $2\pi$. Equation (\ref{match}) is a very suggestive relation, holographic in nature, in which the bare theta angle is obtained in a `mean-field' picture from the parameters of a kind of `instanton condensate'. We should stress that (\ref{match}) is only valid in the non-supersymmetric case. The extremal (supersymmetric) solution has a non-contractible ${\bf S}^1$ so that we can add an arbitrary harmonic piece to $C_{\rm D0}$, thereby changing the asymptotic value of $\theta$ independently of $k$ and $\beta$ (\mbox{{\em i.e.~}} we cannot use Stokes's theorem as we do in (\ref{thetaf}) and (\ref{thotra})). An interesting application of this connection is the computation of topological charge correlations to the leading order in the large-$N$ and large $\lambda$ limit. In view of (\ref{match}), this can be done by studying the $\kappa$-dependence of the vacuum energy of the ${\rm YM}_4$ theory (or equivalently the thermal free energy of the ${\rm SYM}_5$ theory.) For example, the action can be calculated as $ I= \beta \,E_{\rm YM} - S_{\rm BH}$, with $S_{\rm BH}=(A_{\rm horizon})/4G_{10}$ the black-brane entropy and $E_{\rm YM} = M_{\rm ADM} - N\,V \,T_{\rm D4}$, the ADM mass above extremality. One obtains \begin{equation} \label{ac} I={3\,{\rm Vol} ({\bf S}^4) \,\beta V \over 16\pi G_{10}} \,r_0^3 \left(H_0 (r_0) - {7\over 6}\right) = N^2 \,{4VT \over \pi \lambda^2} \,u_0^3 \left( H_0 (u_0) - {7\over 6}\right) .\end{equation} Solving $\theta$ from (\ref{match}) and using the relation \begin{equation} \left({u_k \over u_0}\right)^3 = {6\pi^3 \over c} \,H_0 (u_0) \cdot {\lambda \,\theta \over N} ,\end{equation} combined with (\ref{acheo}) and (\ref{hatt}), we learn that the functional form of the $n=0$ vacuum energy is given by \begin{equation} \label{ff} I(\theta)_{n=0} = N^2 \,VT^4 \, \lambda \,f(\lambda \theta/N) ,\end{equation} with $f(x)$ an even function (as expected from considerations of CP symmetry), whose Taylor expansion around $\theta=0$ may be determined by solving (\ref{hatt}) iteratively. These selection rules determine the large-$N$ and large $\lambda$ scaling of the topological charge correlators at $\theta=0$: \begin{equation} \left\langle \;(Q_{\rm top})^{2m} \;\right\rangle^{\theta=0}_{\rm connected} \,=\,\left({d \over d\theta}\right)_{\theta=0}^{2m} \,I(\theta)\,\sim \,VT^4\; {\lambda^{2m+1}\over N^{2m-2}} .\end{equation} For the standard topological susceptibility, $m=1$, the scaling agrees with ref.\ \cite{hashi}. \section{Concluding Remarks} Within the AdS/CFT correspondence, the large-$N$ master field of the gauge theory is encoded in the gravitational saddle-points of the supergravity description, subject to boundary conditions. In the model of ref. \cite{wit2}, which has a good supergravity description for large $N$ and large 't Hooft coupling $\lambda=g^2 N$, there are two `master fields', or generalized large-$N$ saddle-points, given by the two manifolds ${\bf X}_{\rm vac}$ and ${\bf X}_{\rm bb}$, with ${\bf S}^1 \times {\bf R}^4$ boundary. We find that ${\bf X}_{\rm vac}$ supports instantons in the form of wrapped D0-branes, and leads to exponentially suppressed theta-angle dependence, very much like in the ${\bf AdS}_5 \times {\bf S}^5$ case, to which it is dual through a set of T-duality and localization transitions that we discuss in some detail, including the matching of the single-instanton measure. However, ${\bf X}_{\rm vac}$ is only the dominant master field in the supersymmetric case. The large-$N$ dynamics in the non-supersymmetric case is dominated by ${\bf X}_{\rm bb}$, which {\it does not} support finite-action topological excitations with the instanton charge. Therefore, the dominant master field shows perturbative (in $1/N$) theta-angle dependence, but has no `instanton topology', very much like in the two-dimensional toy models of refs. \cite{toym}. Instead, we can identify approximate (constrained) instantons of size $\rho \ll \beta$, merging with the vacuum at sizes of the order of the glueball's Compton wave-length $\rho \sim \beta$, which for this model coincides with the Kaluza--Klein threshold. The approximate equivalence of ${\bf X}_{\rm vac}$ and ${\bf X}_{\rm bb}$ in the ultraviolet regime $u\rightarrow \infty$, poses the question of whether the approximate small instantons of ${\bf X}_{\rm bb}$ are really artifacts of the regularization of the Yang--Mills theory by a hot five-dimensional supersymmetric theory. Unfortunately, this question cannot be settled with present techniques, since $M_{\rm glue} \sim M_{\rm KK}$ in the supergravity approximation, $\lambda \gg 1$, and we lack a regime in which we could follow the instantons as genuine four-dimensional configurations. It would be very interesting to see if the non-supersymmetric gravity duals based on Type 0 D-branes \cite{tipocero} provide a more vantageous point to study this question. Heuristically, according to the UV/IR relation, an instanton of size $\rho \gg \beta$ would be associated to a D0-brane `inside' the horizon of the black D4-brane. Because of the no-hair properties, such a configuration would have the D0-charge completely de-localized over the horizon (see \cite{mpeet} for a recent discussion in the extremal case). Therefore, such configurations should be interpreted as homogeneous self-dual backgrounds in the gauge theory, and the supergravity description involves the `smeared' D0/D4 solution. Although this picture cannot be held literally in the euclidean solutions, which lack an `interior region' behind the horizon, we can still identify the RR two-form generated by the D0-branes `dissolved' in the D4-brane horizon. This RR flux is in turn responsible for the generation of a theta angle, via the AdS/CFT rules of \cite{gkpw}. Therefore, we obtain a holographic relation between the theta angle and the smeared instanton charge. Although the general relation between background fields and theta angle is not new (see \cite{cole, toym} for explicit two-dimensional examples), we find it interesting that in our case the background field is explicitly associated to an instanton condensate, with quantized topological charge (equal to the number $k$ of smeared D0-branes). This is reminiscent of the instanton liquid models, where the instanton density is fixed self-consistently (see for instance \cite{shu}). \section*{Acknowledgements} We would like to thank Margarita Garc\'{\i}a P\'erez, Yaron Oz and Kostas Skenderis for useful discussions. This work is partially supported by the European Commission TMR programme ERBFMRX-CT96-0045 in which J.L.F.B.\ is associated to the University of Utrecht and A.P.\ is associated to the Physics Department, University of Milano. A.P.\ would like to thank CERN for its hospitality while part of this work was carried out.
\section{Introduction} The {two-dimensional} {one-component} plasma (2dOCP) is a model in classical statistical mechanics which consists of $N$ mobile point particles of charge $q$ interacting on a surface with uniform neutralizing background charge density. The pair potential $\Phi(\vec{r},\vec{r}\,')$ between particles is the solution of the Poisson equation on the particular surface. In the plane \begin{equation}\label{1.1} \Phi(\vec{r},\vec{r}\,') = -\log\Big (|\vec{r} - \vec{r}\,'|/l \Big), \label{1.1ab} \end{equation} where $l$ is some arbitrary length scale which will henceforth be set to unity. With the potential (\ref{1.1}) and a uniform background of charge density $-\rho_b$ inside a disk of radius $R$ $(\rho_b = N/\pi R^2)$ the corresponding Boltzmann factor, which consists of the particle-particle, particle-background and background-background interaction, is given by \begin{equation}\label{1.2} e^{-\Gamma N^2 ( (1/2)\log R - 3/8 )} e^{- \pi \Gamma \rho_b \sum_{j=1}^N |\vec{r}_j|^2/2} \prod_{1 \le j <k \le N} |\vec{r}_k - \vec{r}_j|^\Gamma, \end{equation} where $\Gamma := q^2/k_BT$ is the coupling. We remark that with $\Gamma/2$ an odd integer, (\ref{1.2}) is proportional to the absolute value squared of the celebrated Laughlin trial wave function for the fractional quantum Hall effect \cite{La83}. At the analytic level our knowledge of the properties of the 2dOCP comes from two main sources. First, for the special coupling $\Gamma = 2$, the exact free energy and correlation functions can be calculated for a number of different geometries \cite{AJ81,Ch81,Ca81,JT98}. Second, the 2dOCP is an example of a Coulomb system in its conductive phase and as such should obey a number of sum rules (see e.g.~\cite{Ma88}) which typically represent universal properties of such a system. We remark also that the exact solutions at $\Gamma = 2$ have been an important source of inspiration to identify universal properties. In this paper we develop exact numerical solutions at the special couplings $\Gamma = 4$ and $\Gamma = 6$ for values of $N$ up to 11 and 9 respectively. By undertaking this study we are able to test the prediction of Jancovici et al.~\cite{JMP94} that the expression for the free energy $F$ as a function of the number of particles $N$ be of the form \begin{equation}\label{1.F} \beta F = A N + B N^{1/2} + {\chi \over 12} \log N + \cdots, \end{equation} where $\chi$ denotes the Euler characteristic of the surface ($\chi = 1$ for a disk, $\chi = 2$ for a sphere). Furthermore we are able to investigate the rate of convergence of the one and two point correlation to their thermodynamic values, as well as the accuracy of certain sum rules in the finite system. In fact the latter line of investigation leads us to a new sum rule valid for general $\nu$ dimensional multicomponent Coulomb systems in a spherical domain, which relates to the second moment of the density-charge correlation function in the finite system. We recall (see e.g.~\cite{Ma88}) that in the infinite system the second moment of the charge-charge correlation function is of a universal form known as the Stillinger-Lovett condition. Indeed our sum rule (\ref{eq:thenewsumrule}) below gives the finite size correction to this universal form in systems with a background. As an outline of the paper, we note here that in Section 2 formulas are presented specifying the partition function and one and two point distribution functions for the disk and sphere geometries, with the coupling an even integer, in terms of certain expansion coefficients. These expansion coefficients are in general computationally expensive, but reasonably efficient algorithms exist in the literature applicable to the cases $\Gamma = 4$ and 6. Our numerical results our presented in Section 3. The new sum rules are derived and discussed in Section 4, while Section 5 concludes with a summary. \section{Formalism}\label{sec:formalism} \setcounter{equation}{0} Our interest is in the exact numerical computation of the partition function and one and two-point correlation functions for the 2dOCP in a disk and on the surface of a sphere. In the former system the Boltzmann factor is given by (\ref{1.1}). Two versions of this model will be considered: one in which the particles are confined to a disk of radius $R$ (the same disk which contains the smeared out neutralizing background), and the other in which the particles are can move throughout the plane. These will be referred to as the hard disk and soft disk respectively. In the latter system the Boltzmann factor (\ref{1.1}) is assumed valid also for $|\vec{r}_i| \ge R$, even though the one body potential $\pi \rho_b |\vec{r}_i|^2/2$ is not the correct potential for the coupling between a particle and the background in this region (according to Newton's theorem outside the disk the background creates the same potential as a charge $-N$ at the origin, so the correct Coulomb potential outside the disk is $N \log |\vec{r}_i|$). On the surface of the sphere the Boltzmann factor is given by \begin{equation}\label{2.1} {\left({1\over{2R}}\right)}^{N\Gamma/2}e^{\Gamma N^2/4}\prod_{1\leq j<k\leq N}{|u_kv_j-u_jv_k|}^\Gamma, \end{equation} where $u:=\cos(\theta/2)e^{i\phi/2},\ v:=-i\sin(\theta/2)e^{-i\phi/2}$ are the Cayley-Klein parameters and $(\theta,\phi)$ are the usual spherical coordinates. For our purpose it is convenient to consider the stereographic projection of this system from the south pole of the sphere to the plane tangent to the north pole. This is specified by the equation \begin{equation}\label{2.2} z = 2R e^{i \phi} \tan {\theta \over 2}, \qquad z = x + i y. \end{equation} We then have \begin{eqnarray}\label{2.3}\lefteqn{ {\left({1\over{2R}}\right)}^{N\Gamma/2}e^{\Gamma N^2/4}\prod_{1\leq j<k\leq N}{|u_kv_j-u_jv_k|}^\Gamma dS_1 \cdots dS_N = {\left({1\over{2R}}\right)}^{N\Gamma/2}} \nonumber \\&& \times e^{\Gamma N^2/4} \prod_{j=1}^N {1 \over (1 + |z_j|^2/(4R^2))^{2 + \Gamma (N-1)/2}} \prod_{1\leq j<k\leq N} \left| \frac{z_j - z_k}{2R} \right|^\Gamma d\vec{r}_1 \cdots d\vec{r}_N. \end{eqnarray} \subsection{The cases $\Gamma = 4p$} For $\Gamma = 4p$, integrals over the Boltzmann factors (\ref{1.1}) and (\ref{2.3}) can be performed from knowledge of the coefficients in the expansion \begin{equation}\label{2.S1} \prod_{1 \le j < k \le N}(z_k - z_j)^{2p} = \sum_\mu c_\mu^{(N)}(2p) m_\mu(z_1,\dots,z_N) \end{equation} where $\mu = (\mu_1,\dots,\mu_N)$ is a partition of $pN(N-1)$ such that $$ 2p(N-1) \ge \mu_1 \ge \cdots \mu_N \ge 0 $$ and $$ m_\mu(z_1,\dots,z_N) = {1 \over \prod_i m_i!} \sum_{\sigma \in S_N} z_{\sigma(1)}^{\mu_1}\cdots z_{\sigma(N)}^{\mu_N} $$ is the corresponding monomial symmetric function (the $m_i$ denote the frequency of the integer $i$ in the partition). The key point for the utility of (\ref{2.S1}) is that with $z_j = r_j e^{i \theta_j}$, the $m_\mu$ are orthogonal with respect to angular integrations: \begin{eqnarray}\label{15.o1} \int_0^\infty dr_1 \, r_1 g(r_1^2) \cdots \int_0^\infty dr_N \, r_N g(r_N^2) \int_0^{2 \pi} d \theta_1 \cdots \int_0^{2 \pi} d\theta_N \, m_\mu(z_1,\dots,z_N) \overline{m_\kappa(z_1,\dots,z_N)} \nonumber \\ = \delta_{\mu,\kappa} {N! \over \prod_i m_i!} \pi^N \prod_{l=1}^N G_{\mu_l} \hspace{2cm} \end{eqnarray} where $ G_{\mu_l} := 2 \int_0^\infty dr \, r^{1 + 2\mu_l} g(r^2)$ for arbitrary $g(r^2)$. Thus, after also noting that \begin{equation}\label{2.PD} \prod_{j < k}|z_k - z_j|^{4p} = \prod_{j < k}(z_k - z_j)^{2p} \prod_{j < k}(\bar{z}_k - \bar{z}_j)^{2p}, \end{equation} we see that for $\Gamma = 4p$ \begin{equation}\label{2.A*} I_{N,\Gamma}[g] := \int_{{\mathbf{R}}^2}d\vec{r}_1 \, g(r_1^2) \cdots \int_{{\mathbf{R}}^2}d\vec{r}_N \, g(r_N^2) \, \prod_{j < k} |\vec{r}_k - \vec{r}_j|^{\Gamma} = N! \pi^N \sum_\mu {(c_\mu^{(N)}(2p))^2 \over \prod_i m_i!} \prod_{l=1}^N G_{\mu_l}. \end{equation} In the case $p=1$ this formalism has been utilized by Samaj et al.~\cite{SPK94}, who furthermore presented an algorithm for the computation of $\{c_\mu\}$ in this case. Let us now consider this latter point. In general the coefficients $c_\mu^{(N)}(2p)$ can be calculated from the formula \begin{equation}\label{2.t} c_\mu^{(N)}(2p) = {1 \over (2 \pi)^N} \int_0^{2 \pi} d\theta_1 \, e^{-i \mu_1 \theta_1} \cdots \int_0^{2 \pi} d\theta_N e^{-i\mu_N \theta_N} \prod_{j < k} (e^{i \theta_k} - e^{i \theta_j})^{2p}, \end{equation} which follows from (\ref{2.S1}). Since we require $|\mu| = pN(N-1)$, the integral over $\theta_N$ can be performed by changing variables $\theta_j \mapsto \theta_j + \theta_N$ $(j=1,\dots,N-1)$ to give \begin{eqnarray}\label{2.tt} c_\mu^{(N)}(2p) & = & {1 \over (2 \pi)^{N-1}} \int_0^{2 \pi} d\theta_1 \, e^{-i \mu_1 \theta_1} \cdots \int_0^{2 \pi} d\theta_{N-1} e^{-i\mu_{N-1} \theta_{N-1}} \prod_{j=1}^{N-1}(1 - e^{i\theta_j})^{2p} \nonumber \\ && \times \prod_{1 \le j < k \le N - 1} (e^{i \theta_k} - e^{i \theta_j})^{2p}. \end{eqnarray} The simplest case is $N=2$, when the sum over pairs in (\ref{2.tt}) is not present. Expanding $(1 - e^{i\theta_1})^{2p}$ according to the binomial theorem gives $$ c_\mu^{(2)}(2p) = (-1)^{\mu_1} \left ( {2p \atop \mu_1} \right ) $$ where $\mu_1 = p,p+1,\dots,2p$ (for $\mu_1 = p$ we have $\mu_1 = \mu_2$ and thus $m_{\mu_1} = 2$, while in all other cases $\mu_1 \ne \mu_2$ and so $m_{\mu_1} = m_{\mu_2} = 1$). Substituting in (\ref{2.A*}) we see, after some minor manipulation, that \begin{equation}\label{2.2'} \int_{{\mathbf{R}}^2}d\vec{r}_1 \, g(r_1^2) \int_{{\mathbf{R}}^2}d\vec{r}_2 \, g(r_2^2) \, |\vec{r}_2 - \vec{r}_1|^{4p} = \pi^2 \sum_{\mu=0}^{2p} \left ( {2p \atop \mu} \right )^2 \int_0^\infty dr \, r^\mu g(r) \int_0^\infty dr \, r^{2p - \mu} g(r). \end{equation} To calculate $c_\mu^{(N)}(2p)$ via this method for a general value of $N$ would require expanding ${1 \over 2}(N-1)N$ products via the binomial theorem, giving a total of $({1 \over 2}(N-1)N)^{2p+1}$ terms to determine each value of $c_\mu$. Thus for a given value of $N$ the complexity increases exponentially with the coupling $p$. As we want to determine the $c_\mu$ for a sequence of values of $N$ as large as possible, we are therefore restricted to the case $p=1$. In fact the case $p=1$ allows (\ref{2.t}) to be computed without using the binomial expansion \cite{SPK94}. Instead one uses the Vandermonde formula for the product of differences as a determinant to expand the products in (\ref{2.t}). This gives \begin{equation}\label{2.van} c_\mu^{(N)}(2) = \sum_{P \in S_N} \varepsilon(P) \sum_{Q \in S_N} \varepsilon(Q) \prod_{k=1}^N \delta_{P(k) + Q(k) - 2,\mu_k} = \sum_{P \in S_N} \varepsilon(P) \sum_{Q \in S_N} \prod_{k=1}^N \delta_{P(k) + k - 2,\mu_{Q(k)}}, \end{equation} which is the formula we used to compute our data in the case $p=1$ for $N=3,\dots,10$. \subsection{The cases $\Gamma = 4p+2$} With $\Gamma = 4p+2$, decomposing the product of differences analogous to (\ref{2.PD}) shows that we must consider the product of differences raised to an odd power. The analogue of (\ref{2.S1}) is then the expansion \begin{equation}\label{15.1'} \prod_{1 \le j < k \le N} (z_k - z_j)^{2p+1} = \sum_\mu c_\mu^{(N)}(2p+1) {\cal A}(z_1^{\mu_1 + N - 1}z_2^{\mu_2 + N - 2} \cdots z_N^{\mu_N}) \end{equation} where $2p(N-1) \ge \mu_1 \ge \mu_2 \ge \cdots \ge \mu_N \ge 0$, $\sum_{j=1}^N \mu_j = p N (N -1)$ and ${\cal A}$ denotes antisymmetrization. Factoring out the antisymmetric factor $\prod_{j < k}(z_k - z_j)$ from both sides then gives \begin{equation}\label{15.2'} \prod_{1 \le j < k \le N} (z_k - z_j)^{2p} = \sum_\mu c_\mu^{(N)}(2p+1) S_\mu(z_1,\dots,z_N) \end{equation} where $S_\mu$ denotes the Schur polynomial indexed by the partition $\mu$. Furthermore, analogous to the orthogonality (\ref{15.o1}) we have \begin{eqnarray}\label{15.o2} \int_0^\infty\cdots\int_0^\infty \prod_{l=1}^N dr_l\, r_l g(r_l^2) \int_0^{2 \pi}d\theta_1 \cdots \int_0^{2 \pi} d\theta_N \prod_{j<k}\left|z_j-z_k\right|^2 S_\mu(z_1,\dots,z_N) \overline{S_\kappa(z_1,\dots,z_N)} \nonumber \\ = \delta_{\mu,\kappa} N!\pi^N \prod_{l=1}^N G_{\mu_l+N-l}. \hspace{2cm} \end{eqnarray} Thus for $\Gamma = 4p+2$, instead of (\ref{2.A*}) we have \begin{equation}\label{2.B*} I_{N,\Gamma}[g] = N! \pi^N \sum_\mu (c_\mu^{(N)}(2p+1))^2 \prod_{l=1}^N G_{\mu_l+N-l}. \end{equation} According to (\ref{15.1'}) the coefficients $c_\mu^{(N)}(2p+1)$ can be computed from the formula (\ref{2.t}) with $\mu_j \mapsto \mu_j + N - j$ and $2p \mapsto 2p+1$, or equivalently (\ref{2.tt}) with the same replacements. In the case $N=2$ this latter formula gives $$ c_\mu^{(2)}(2p+1) = (-1)^{\mu_1 + 1} \left ( {2p +1 \atop \mu_1 + 1} \right ) $$ with $\mu_1 = p,\dots, 2p$. This in turn implies that the formula (\ref{2.2'}) again holds with $2p \mapsto 2p+1$. To obtain data for consecutive values of $N$, the computationally simplest case is $p=1$. However algorithms based on (\ref{2.t}) (with $\mu_j \mapsto \mu_j + N - j$ and $2p \mapsto 2p+1$) are inferior to methods that determine $c_\mu^{(N)}(3)$ from (\ref{15.2'}) \cite{DGIL94,Du94,STW94}. The most efficient algorithm appears to be the one of Scharf et al.~\cite{STW94}, where the coefficients $c_\mu^{(N)}(3)$ are determined up to $N=9$. Fortunately the authors of \cite{STW94} have kindly supplied us with their data (up to $N=8$), so we do not need to repeat the calculation. \subsection{The sphere} The Boltzmann factor for the sphere, stereographically projected onto the plane, is given by the r.h.s.~of (\ref{2.3}). Thus, with $\r_j\mapsto2R\r_j$ we require \begin{equation}\label{2.g1} g(r^2) = (1 + r^2)^{-(N-1)\Gamma/2 - 2} \end{equation} in the integral (\ref{2.A*}). However, computational savings can be obtained by first noting that because the sphere is homogeneous, one particle can be fixed at the north pole, reducing the number of integrals from $N$ to $N-1$ (we must also multiply by $\pi$ -- the area of the surface of a sphere of radius $1/2$). Thus we have \begin{eqnarray} \int_{({\mathbf{R}}^2)^N} d\vec{r}_1 \cdots d\vec{r}_N \, \prod_{i=1}^N {1 \over (1 + |z_i|^2)^{(N-1)\Gamma/2 + 2}} \prod_{1 \le j < k \le N} |z_k - z_j|^\Gamma \qquad \nonumber \\ = \pi \int_{({\mathbf{R}}^2)^{N-1}} d\vec{r}_1 \cdots d\vec{r}_{N-1} \, \prod_{i=1}^{N-1} {|z_i|^\Gamma \over (1 + |z_i|^2)^{(N-1)\Gamma/2 + 2}} \prod_{1 \le j < k \le N-1} |z_k - z_j|^\Gamma, \end{eqnarray} and so should choose \begin{equation}\label{2.g2} g(r^2) = {r^\Gamma \over (1 + r^2)^{(N-1)\Gamma / 2 + 2} } \end{equation} in (\ref{2.A*}). With $g(r^2)$ given by (\ref{2.g2}), the formulas (\ref{2.A*}) and (\ref{2.B*}) show that at $\Gamma = 4$ and $\Gamma = 6$ the canonical partition function $$ Z_{N,\Gamma} := {1 \over N!} \int_{({\mathbf{R}}^2)^{N}} d\vec{r}_1 \cdots d\vec{r}_N \, e^{-\beta U} $$ can be represented by the series \begin{eqnarray} Z_{N+1,4}^{\mathrm{sphere}} & = & {e^{(N+1)^2} \pi^{N+1} \over N + 1} \sum_\nu \left(c_\nu^{(N)}(2)\right)^2 {1 \over \prod_i m_i!} \prod_{i=1}^N {(\mu_i + 2)!(2N - \mu_i - 2)! \over (2N+1)!} \label{2.s1}\\ Z_{N+1,6}^{\mathrm{sphere}} & = & \rho_b^{(N+1)/2} (N+1)^{(N+3)/2} e^{3(N+1)^2/2} \pi^{3(N+1)/2} \sum_\nu (c_\nu^{(N)}(3))^2 \nonumber \\ && \times \prod_{k=1}^N {(3+N + \mu_k - k)! (2N - 3 - \mu_k + k)! \over (3N + 1)!}. \label{2.s2} \end{eqnarray} To obtain these formulas use has been made of the definite integral \begin{equation}\label{2.defi} \int_0^\infty {r^p \over (1 + r)^q} \, dr = {\Gamma(p+1) \Gamma (q - p - 1) \over \Gamma (q) }. \end{equation} Because the sphere is homogeneous, the two-point distribution $\rho_{(2)}((\theta,\phi), (\theta',\phi'))$ can be computed with one particle at the north pole ($\theta' = 0$ say). We then have $$\rho_{(2)}((\theta,\phi), \, (\theta',\phi')) = \rho_{(2)}(\theta) $$ so the two-point function can be computed from an integral of the form (\ref{2.A*}). In fact with $g(r^2)$ given by (\ref{2.g1}) we have \begin{equation} \rho_{(2)}(\theta) = \frac{1}{4R^2} \frac{1}{I_{N,\Gamma}[g]}\lim_{x'\to0} \frac{g(x^2)g({x'}^2)}{4\pi^2 xx'} (1+x^2)^2(1+x'^2)^2 \frac{\delta^2 I_{N,\Gamma}[g]}{\delta g(x^2) \delta g(x'^2)} \end{equation} where $x=\tan(\theta/2)$. For $\Gamma = 4$ this gives \begin{eqnarray} \lefteqn{\rho_{(2)}(\theta) =\rho_b^2\frac{(2N-1)!}{N^2(1+x^2)^{2N-2}}} \nonumber\\ &\times&\frac{\displaystyle \sum_{\mu,\mu_N=0}(c_\mu^{(N)}(2))^2 {1 \over \prod_i m_i!} \prod_{i=1}^N \mu_i!(2N-2-\mu_i)! \sum_{k=1}^{N-1} \frac{x^{2\mu_k}}{\mu_k!(2N-2-\mu_k)!} }{ \sum_{\mu}(c_\mu^{(N)}(2))^2 {1 \over \prod_i m_i!} \prod_{i=1}^N \mu_i!(2N-2-\mu_i)! } \end{eqnarray} while for $\Gamma = 6$ we deduce that \begin{eqnarray} \lefteqn{\rho_{(2)}(\theta) =\rho_b^2\frac{(3N-2)!}{N^2(1+x^2)^{3N-3}} \,(3N-2) } \nonumber\\ &\times&\frac{ \displaystyle \sum_{\mu,\mu_N=0} \!\!(c_\mu^{(N)}(3))^2 \! \prod_{i=1}^N\! (\mu_i\!\!+\!\!N\!\!-\!\!i)!(2N\!\!-\!\!3\!\!-\!\!\mu_i\!\!+\!\!i)! \!\sum_{k=1}^{N-1} \frac{x^{2(\mu_k+N-k)}}% {(\mu_k\!\!+\!\!N\!\!-\!\!k)!(2N\!\!-\!\!3\!\!-\!\!\mu_k\!\!+\!\!k)!}}{ \sum_{\mu}(c_\mu^{(N)}(3))^2 \prod_{i=1}^N (\mu_i+N-i)!(2N-3-\mu_i+i)! } \end{eqnarray} \subsection{The disk} In the case of the disk, (\ref{1.2}) with $\vec{r}_j \mapsto R \vec{r}_j$ shows we require \begin{equation}\label{2.g2'} g(r^2) = \chi(r) e^{-\Gamma N |\vec{r}_j|^2/2} \end{equation} where $\chi=1$ for $r^2 < 1$ and zero otherwise in the case of the hard disk, while $\chi=1$ for all $r$ in the case of the soft disk. Thus from (\ref{2.A*}) we have at $\Gamma = 4$ \begin{eqnarray} Z_{N,4}^{\mathrm{soft\ disk}}& = & e^{3N^2/2}\left(\frac{1}{2N}\right)^{N^2} \pi^N \sum_\mu (c_\mu^{(N)}(2))^2 \left(\prod_i m_i!\right)^{-1} \prod_{i=1}^N \mu_i! \label{3.sa}\\ Z_{N,4}^{\mathrm{hard\ disk}}& = & e^{3N^2/2}\left(\frac{1}{2N}\right)^{N^2} \pi^N \sum_\mu (c_\mu^{(N)}(2))^2 {1 \over \prod_i m_i !} \prod_{i=1}^N \gamma(\mu_i+1,2N) \label{3.sb} \end{eqnarray} while at $\Gamma = 6$ use of (\ref{2.B*}) gives \begin{equation} \label{3.sc} Z_{N,6}^{\mathrm{hard}} = \rho_b^{N/2}N^{-3N^2/2}3^{-N(3N-1)/2} \pi^{3N/2}e^{9N^2/4} \sum_\mu (c_\mu^{(N)}(3))^2 \prod_{k=1}^N \gamma(\mu_k+N-k+1,3N) \end{equation} with the soft disk case obtained by replacing the incomplete gamma functions by complete gamma functions. Unlike the situation with the sphere, the density is a non-constant function in the disk geometry. Now, with $g(r^2)$ given by (\ref{2.g2}) we have $$ \rho_{(1)}(r) = {g(r^2) \over 2 \pi r} {\delta \log Z_{N,4}^{\mathrm{disk}} \over \delta g(r^2)}. $$ At $\Gamma = 4$ this gives \begin{eqnarray}\label{2.den} \lefteqn{ \rho_{(1)}(r)=2\rho_b e^{-2\pi \rho_b r^2} }&& \nonumber\\ &&\times \frac{\displaystyle \sum_\mu (c_\mu^{(N)}(2))^2 { 1 \over \prod_i m_i!} \prod_{j=1}^N \gamma(\mu_j+1,2N) \sum_{k=1}^N \frac{(2\pi \rho_b r^2)^{\mu_k}}% {\gamma(\mu_k+1,2N)} }{ \sum_\mu (c_\mu^{(N)}(2))^2 { 1 \over \prod_i m_i!} \prod_{j=1}^N \gamma(\mu_j+1,2N) } \end{eqnarray} while at $\Gamma = 6$ one obtains \begin{eqnarray} \lefteqn{ \rho_{(1)}(r)=3\rho_b e^{-3\pi \rho_b r^2} }&& \nonumber\\ &&\times \frac{\displaystyle \sum_\mu (c_\mu^{(N)}(3))^2 \prod_{j=1}^N \gamma(\mu_j+N-j+1,3N) \sum_{k=1}^N \frac{(3\pi \rho_b r^2)^{\mu_k+N-k}}% {\gamma(\mu_k+N-k+1,3N)} }{ \sum_\mu (c_\mu^{(N)}(3))^2 \prod_{j=1}^N \gamma(\mu_j+N-j+1,3N) } \end{eqnarray} The corresponding formulas for the soft disk are obtained by replacing the incomplete gamma functions by complete gamma functions. Finally, we consider the two-point function in the disk geometry. In general this quantity is not just a function of the distance between particles, and so we cannot use the formalism based on the orthogonalities (\ref{15.o1}) and (\ref{15.o2}). However, with one of the particles fixed at the origin ($\vec{r}\,' = \vec{0}$ say) we have $\rho_{(2)}(\vec{r}, \vec{r}\,') = \rho_{(2)}(r)$, so in this case the formalism used to compute the densities can again be used. Thus using the general formula $$ \rho_{(2)}(r) = {1 \over Z_{N,\Gamma}} \lim_{r' \to 0} {g(r^2) g({r'}^2) \over 4 \pi r r'} {\delta^2 Z_{N,\Gamma} \over \delta g(r^2) \delta g({r'}^2)}, $$ we find for the hard disk case \begin{eqnarray}\label{eq:2.rho2-g4} \lefteqn{ \rho_{(2)}(r) = 4 \rho_b^2 e^{-2\pi \rho_b r^2} }&& \nonumber\\ &&\times \frac{ \sum_{\mu,\mu_N=0} (c_\mu^{(N)}(2))^2 {1 \over \prod_i m_i !} \prod_{j=1}^{N-1} \gamma(\mu_j+1,2N) \sum_{k=1}^{N-1} \frac{(2\pi \rho_b r^2)^{\mu_k}}{\gamma(\mu_k+1,2N)} }{ \sum_{\mu} (c_\mu^{(N)}(2))^2 {1 \over \prod_i m_i !} \prod_{j=1}^{N} \gamma(\mu_j+1,2N) } \end{eqnarray} \begin{eqnarray}\label{eq:2.rho2-g6} \lefteqn{ \rho_{(2)}(r) = 9\rho_b^2 e^{-3\pi \rho_b r^2} }&& \nonumber\\ &&\times \frac{\displaystyle \sum_{\mu,\mu_N=0} (c_\mu^{(N)}(3))^2 \prod_{j=1}^{N-1} \gamma(\mu_j+N-j+1,3N) \sum_{k=1}^{N-1} \frac{(3\pi \rho_b r^2)^{\mu_k+N-k}}{\gamma(\mu_k+N-k+1,3N)} }{ \sum_{\mu} (c_\mu^{(N)}(3))^2 \prod_{j=1}^{N} \gamma(\mu_j+N-j+1,3N) } \end{eqnarray} for $\Gamma = 4$ and $\Gamma = 6$ respectively. Again the corresponding results for the soft disk are obtained by replacing the incomplete gamma functions by complete gamma functions. \section{Numerical results} \setcounter{equation}{0} \subsection{Free energy -- sphere geometry} In the Introduction it was commented that the free energy is expected to have a large $N$ expansion of the form (\ref{1.F}) with $\chi = 2$ in sphere geometry. In fact the constant $B$ in (\ref{1.F}), which is a surface free energy, should be identically zero in sphere geometry, so we expect a large $N$ expansion of the form \begin{equation}\label{3.F1} \beta F = A N + {1 \over 6} \log N + C + \cdots. \end{equation} As noted by Jancovici et al.~\cite{JMP94}, the validity of (\ref{3.F1}) can be explicitly demonstrated at $\beta = 2$ because of an exact solution due to Caillol \cite{Ca81}. The mechanism for the exact solution can be seen within the present formalism. Thus, at $\Gamma = 2$ we require the coefficients $c_\mu^{(N)}(1)$ in (\ref{15.1'}). But this follows from the Vandermonde expansion (recall (\ref{2.van})), which gives $c_\mu^{(N)}(1)=1$ for $\mu = 0^N$ and $c_\mu^{(N)}(1) = 0$ otherwise. Substituting in (\ref{2.B*}) with $g(r^2)$ given by (\ref{2.g2}), and making use of (\ref{2.defi}) we thus obtain \cite{Ca81} \begin{equation}\label{3.ca} Z_{N,2}^{\mathrm{sphere}} = \pi^{-N/2} N^{N/2} \rho_b^{-N/2} e^{N^2/2} \prod_{k=1}^N {(N-k)! (k-1)! \over N!}. \end{equation} This substituted into the general formula \begin{equation}\label{3.fr} \beta F_{N, \Gamma} = - \log Z_{N, \Gamma} \end{equation} leads to the expansion \cite{JMP94} \begin{equation}\label{3.fr1} \beta F = Nf + {1 \over 6} \log N + {1 \over 12} - 2 \zeta'(-1) + {\mathrm{o}}(1) \end{equation} where $f = {1 \over 2} \log (\rho_b / 2 \pi^2)$. We remark that by introducing the Barnes $G$ function according to $$ G(z+1) = \Gamma(z) G(z), \quad G(1) = 1 $$ we can write $$ \prod_{k=1}^N (k-1)! = G(N + 1). $$ The large $N$ expansion of the Barnes $G$ function is known to be \cite{Ba00} \begin{equation}\label{3.barnes} G(N+1) \sim {N^2 \over 2} \log N - {3 \over 4} N^2 + {N \over 2} \log 2 \pi - {1 \over 12} \log N + \zeta'(-1) - {1 \over 720 N^2} + {\mathrm{O}}\Big ( {1 \over N^4} \Big ). \end{equation} This together with Stirling's formula allows us to extend (\ref{3.fr1}) to the expansion \begin{equation}\label{3.fr2} \beta F = Nf + {1 \over 6} \log N + {1 \over 12} - 2 \zeta'(-1) + {1 \over 180 N^2} + {\mathrm{O}}\Big ({1 \over N^4} \Big ). \end{equation} In the cases $\Gamma = 4$ and $\Gamma = 6$, by following the numerical procedure detailed in the previous section, we have been able to compute the partition functions (\ref{2.s1}) and (\ref{2.s2}) up to 11 and 9 particles respectively. The results are listed in Table \ref{t3.1}. Our results are presented in decimal form. However the terms in the summations of (\ref{2.s1}) and (\ref{2.s2}) are all rational numbers, and we have also calculated the sum itself as a rational number. A point of interest is the factorization of the denominator and numerator of the rational number. The exact result (\ref{3.ca}) shows that at $\Gamma = 2$ only small integers occur in this factorization. However our exact data shows that this feature is no longer true at $\Gamma = 4$ or $\Gamma = 6$. For example, at $\Gamma = 4$ and with $N=9$ we find that the summation in (\ref{2.s1}) is given by the ratio of primes $$ \frac{19\cdot 23\cdot 31\cdot 404431651134013\cdot 56827}% {2^{28}3^{12}5^7 7^8 11^8 13^8 17^8}. $$ \begin{table} \caption{\label{t3.1} Exact numerical computation of the expressions (\ref{2.s1}) and (\ref{2.s2}) (in the latter case we have set $\rho_b =1$), and the corresponding free energy (\ref{3.fr}). } \vspace{.4cm} {\small \begin{tabular}{l|l|l} $N$ & $Z_{N,4}$ & $\beta F_{N,4}$ \\ \hline 3 & 9.770695753081390794542103296367E+02 & -6.884557862719257767291929292830 \\ 4 & 1.081868103379375397165672403770E+04 & -9.289029644211538110263324038604 \\ 5 & 1.209528877878741526102013133936E+05 & -11.70315639163470461293716934684 \\ 6 & 1.360835037494310939624360869217E+06 & -14.12360906745006986750189927991 \\ 7 & 1.537846289459171693753614603094E+07 & -16.54847857521316551691816164401 \\ 8 & 1.743564157878398325393942744018E+08 & -18.97661212873318180330363390257 \\ 9 & 1.981770773388678655915061613417E+09 & -21.40725661197234419004446417460 \\ 10 & 2.257011016434890100740949944465E+10 & -23.83989230877186989649422160272 \\ 11 & 2.574639922522006241714385546434E+11 & -26.27414571135846506646694529338 \\ \end{tabular} } \vspace{.4cm} {\small \begin{tabular}{l|l|l} $N$ & $Z_{N,6}$ & $\beta F_{N,6}$ \\ \hline 2& 781.80154948970530457541038293910180& -6.661600935308419284761353568226471 \\ 3& 24731.016946702464115291740435512837& -10.115813481655518642906626162676076 \\ 4& 798906.45662411908447403801186279894& -13.590999142330226359670889161470696 \\ 5& 25990836.664099377843271224794515169& -17.073254597869416657276355484106596 \\ 6& 851167572.30792422833993160492670601& -20.562119579383207945093207167461793\\ 7& 27989023411.960800446597844273994987& -24.055078249259894430456119939885817\\ 8& 923260788226.64381072982338145761830& -27.551177575665397081224942401207047 \\ 9& 30529687045074.352434196537904510620& -31.049720671888250916196597607309575 \\ \end{tabular} } \end{table} To analyze our data we first sought fitting sets of consecutive values of $N$ to the ansatz \begin{equation}\label{3.ans} \beta F_{\Gamma, N} = A_\Gamma N + K_\Gamma \log N + C_\Gamma. \end{equation} The results are contained in Table \ref{t3.2}. Notice that at $\Gamma = 4$ the value of the free energy per particle $A$ appears to have converged to 3 decimal place accuracy, while the value of $K$ appears similarly to be converging, with the final value in the table differing from $1/6$ only in the third decimal. The general trends are the same for the $\Gamma = 6$ data, although the convergence rate (as determined by the difference between sequential values) is slower. \begin{table} \caption{\label{t3.2} Fitting the values of $\beta F_{\Gamma, N}$ with $N$ as specified, taken from Table \ref{t3.1}, to the ansatz (\ref{3.ans}). } \vspace{.4cm} \begin{tabular}{l|l|l|lr|l|l|l} $N$ & $A_4$ & $K_4$ & $C_4$ & \qquad & $A_6$ & $K_6$ & $C_6$ \\ \hline 3,4,5& -2.447509 & 0.149600 & 0.293616& & -3.526411 & 0.178065 & 0.267797 \\ 4,5,6& -2.448705& 0.154963 & 0.290968&& -3.506699 & 0.109543 & 0.283938 \\ 5,6,7& -2.449038& 0.156787& 0.289696& & -3.515359& 0.145316 & 0.269664 \\ 6,7,8& -2.449271 & 0.158300 & 0.288384& & -3.516438 & 0.152316 & 0.263596 \\ 7,8,9& -2.449423 & 0.159440 & 0.287231 & & -3.516820 & 0.155176 & 0.260704 \\ 8,9,10 & -2.449524 & 0.160290 & 0.286264 & & & & \\ 9,10,11 & -2.449594 & 0.160960 & 0.285428 & & & & \\ \end{tabular} \end{table} Next we sought fitting four consecutive values of $N$ to the ansatz \begin{equation}\label{3.ans1} \beta F_{\Gamma, N} = A_\Gamma N + K_\Gamma \log N + C_\Gamma +D_\Gamma/N. \end{equation} The results of this fit are presented in Table \ref{t3.3}. At $\Gamma = 4$ this markedly improves the convergence rate, with the final estimate of $K$ now differing from $1/6$ by only 3 parts in $10^4$. However at $\Gamma = 6$ the convergence rate is in fact worsened, indicating some illconditioning when the extra free parameter is introduced. Note also that the coefficient of $1/N$ in both cases appears to be non-zero, as distinct from the situation at $\Gamma = 2$ exhibited by the analytic result (\ref{3.fr2}). \begin{table} \caption{\label{t3.3} Fitting the values of $\beta F_{\Gamma, N}$ with $N$ as specified, taken from Table \ref{t3.1}, to the ansatz (\ref{3.ans1}). } \vspace{.4cm} {\small \begin{tabular}{l|l|l|l|l|lr|l|l|l} $N$ & $A_4$ & $K_4$ & $C_4$ & $D_4$ & \qquad & $A_6$ & $K_6$ & $C_6$ & $D_6$\\ \hline 3,4,5,6 & -2.450743& 0.175200& 0.258672 & 0.049566 & & -3.5382 & 0.3594 & 0.0572& 0.4839 \\ 4,5,6,7 & -2.449773 & 0.165568 & 0.2740449 & 0.025973 && -3.5086& 0.0654 & 0.4121 & 0.2363 \\ 5,6,7,8 & -2.449905 & 0.167146 & 0.2712323 & 0.031065 && -3.5193 & 0.1932 & 0.1842 & 0.1417 \\ 6,7,8,9 & -2.449914 & 0.167268 & 0.2709949 & 0.031065 && -3.5180 & 0.1748 & 0.2199 & 0.0779 \\ 7 ,8,9,10 & -2.449896 & 0.166989 & 0.2715743 & 0.029956 && & & & \\ 8,9,10,11 & -2.449892 & 0.166917 & 0.2717321 & 0.029634 \\ & & & \\ \end{tabular} } \end{table} Finally, we sought to estimate from our data an accurate as possible value of the free energy per particle, $\beta f_\Gamma$ say. For this purpose we fitted the data to the ansatz \begin{equation}\label{3.ans2} \beta F_{\Gamma, N} = A_\Gamma N + {1 \over 6} \log N + C_\Gamma +D_\Gamma/N + \left \{ \begin{array}{ll} E_\Gamma/N^2, & \Gamma = 4 \\ 0, & \Gamma = 6 \end{array} \right. \end{equation} thus assuming the universal term in (\ref{3.F1}). Four free parameters are used at $\Gamma = 4$, while only 3 free parameter are used at $\Gamma = 6$, in keeping with observed illconditioning when a fourth parameter is introduced. Our results are presented in Table \ref{t3.4}, where $\beta f_\Gamma$ is determined by $A_\Gamma$. We see that there at $\Gamma = 4$ we appear to have convergence to 7 digits with the estimate \begin{equation}\label{3.e1} \beta f_4 = -2.449884 \cdots \end{equation} while at $\Gamma = 6$ our final estimate is \begin{equation}\label{3.e2} \beta f_6 = -3.5175 \cdots \end{equation} accurate to 5 digits. \begin{table} \caption{\label{t3.4} Fitting the values of $\beta F_{\Gamma, N}$ with $N$ as specified, taken from Table \ref{t3.1}, to the ansatz (\ref{3.ans2}). } \vspace{.4cm} {\small \begin{tabular}{l|l|l|l|lr|l|l|l|l} $N$ & $A_4$ & $C_4$ & $D_4$ & $E_4$ & \qquad & $A_6$ & $C_6$ & $D_6$ \\ \hline 3,4,5,(6) & -2.4501031 & 0.275576 & 0.012460 & 0.026276 && -3.513916& 0.205966& 0.110598 \\ 4,5,6,(7) & -2.4498406 & 0.271639 & 0.031880 & 0.005215 && -3.518863& 0.250494&0.011648 \\ 5,6,7,(8) & -2.4498809 & 0.272364 & 0.027574 & 0.003235 && -3.517146 & 0.231609 & 0.063153\\ 6,7,8,(9) & -2.4498875 & 0.272503 & 0.026605 & 0.005465 && -3.517466& 0.235770& 0.049709 \\ 7,8,9,(10) & -2.4498842 & 0.272423 & 0.027240 & 0.003788 && -3.517540& 0.236870& 0.045600 \\ 8,9,10,11 & -2.4498841 & 0.272420 & 0.027272 & 0.003695 && & & \\ \end{tabular} } \end{table} We note that there is some early literature on estimating $\beta f_4$ and $\beta f_6$ from exact small $N$ numerical data \cite{JM83}. Using only the values of $\beta F_{N,\Gamma}$ for $N=1,2$ and 3, the quantity $$ \beta \tilde{f}_\Gamma = \beta f_\Gamma + \Big ( {3 \Gamma \over 8} + 1 \Big ) + {\Gamma \over 4} \log \pi \rho_b - \log \rho_b $$ was estimated for $\Gamma = 4,6,\dots,10$. In particular, at $\Gamma = 4$ and $\Gamma = 6$ these estimates of $\beta f_\Gamma$ give $$ \beta f_4 \approx -2.1585, \qquad \beta f_6 \approx -3.330, $$ which differ from our estimates (\ref{3.e1}) and (\ref{3.e2}) in the first decimal place. \subsection{Free energy -- disk geometry} For the disk geometry, the prediction (\ref{1.F}) gives a large $N$ expansion of the form \begin{equation}\label{3.F} \beta F_\Gamma = AN + BN^{1/2} + {1 \over 12} \log N + C + \cdots \end{equation} As in the case of the sphere geometry, this prediction can be verified analytically using the exact solution for the isotherm $\Gamma = 2$ \cite{AJ81}. The exact solution gives \cite{JMP94} \begin{equation}\label{3.F1'} \beta F_2^{\mathrm{hard}} = \beta f_2 N + \beta \gamma_2 N^{1/2} + {1 \over 12} \log N + {\mathrm{O}}(1) \end{equation} where $$ \beta f_2 = {1 \over 2} \log (\rho_b / 2 \pi^2), \quad \beta \gamma_2 = \sqrt{2} \int_0^\infty dy \, \log \Big ( {1 \over 2} (1 + {\mathrm{erf}} \, y) \Big ). $$ Some details of the expansion of $\beta F_2$ are different for the soft edge version of the OCP in a disk (recall Section 1). From the exact formula $$ Z_{N,2}^{\mathrm{soft}} = \pi^N e^{3 N^2 / 4} N^{-N^2/2} (\pi \rho_b )^{-N/2} G(N+1) $$ and the asymptotic expansion (\ref{3.barnes}) we see that \begin{equation}\label{3.F2} \beta F_2^{\mathrm{soft}} = \beta f N + {1 \over 12} \log N - \zeta ' (-1) - {1 \over 720 N^2} + {\mathrm{O}} \Big ( {1 \over N^4} \Big ). \end{equation} Thus indeed both (\ref{3.F1}) and (\ref{3.F2}) contain the universal term $(1/12) \log N$, although (\ref{3.F2}) does not contain a surface tension term (this fact has been noted previously in \cite{DGIL94}). At $\Gamma = 4$ and $\Gamma = 6$ we obtained exact numerical evaluation of the partition functions (\ref{3.sa}), (\ref{3.sb}) and (\ref{3.sc}) (and the modification of (\ref{3.sc}) for the soft disk case) as in the sphere case. Our results for the corresponding value of $\beta F$ are contained in Table \ref{t3.5}. To test the prediction (\ref{3.F1}), we sought to fit our data to the ansatz \begin{equation}\label{3.F3} \beta F_{N,\Gamma} = \beta f_\Gamma N + B_\Gamma N^{1/2} + K_\Gamma \log N + C_\Gamma + \left \{ \begin{array}{ll} D_\Gamma/N, & {\mathrm{soft \, disk}} \\ 0, & {\mathrm{hard \, disk}} \end{array}\right. \end{equation} where $\beta f_\Gamma$ is given by (\ref{3.e1}) and (\ref{3.e2}) for $\Gamma = 4$ and $\Gamma = 6$ respectively, and the choice in (\ref{3.F3}) is made retrospectively on the criterium of obtaining better convergence. Our results are obtained in Table \ref{t3.6}. We see that for the hard disk at $\Gamma = 4$ our final estimate of $K_4$ differs from $1/12$ by only $2 \times 10^{-4}$. At $\Gamma = 6$ we see that more data would be needed to get a stable sequence, although the final estimates of $K_6$ are consistent with the expected value of $1/12$. \begin{table} \caption{\label{t3.5} Exact decimal expansion of the free energy for the hard and soft disk at $\Gamma = 4$ and $\Gamma = 6$ } \vspace{.4cm} {\small \begin{tabular}{l|l|l} $N$ & $F_{N,4}^{\mathrm{hard}}$ & $\beta F_{N,4}^{\mathrm{soft}}$ \\ \hline 3 & -6.07705853011644579848828232852953 & -6.38430353764202167882687100789504 \\ 4 & -8.30894530308837749094468707356467 & -8.67246771929839719253598118439664 \\ 5 & -10.5685824419856069054748395707000 & -10.9817913623032469741300225072724 \\ 6 & -12.8480499008173510151377678768908 & -13.3060582270200975291371029447052 \\ 7 & -15.1423987396644292302500680775824 & -15.6414978836634761215222474874096 \\ 8 & -17.4483520149155330139161065965798 & -17.9856458201720068377211714643235 \\ 9 & -19.7636864904052121059815096874218 & -20.3368227969313363262711724430690 \\ 10 & -22.0868149972503557220763154840028 & -22.6938278975627003536283880871543 \end{tabular} } \vspace{.4cm} \vspace{.4cm} {\small \begin{tabular}{l|l|l} $N$ & $F_{N,6}^{\mathrm{hard}}$ & $\beta F_{N,6}^{\mathrm{soft}}$ \\ \hline 3 & -9.0582041809587470427592556776938317 & -9.1916690110088058948684895153913657 \\ 4 & -12.306265058620940233015626198772823 & -12.467150515773535356614120708869630 \\ 5 & -15.583591141405785588643527765993475 & -15.769625685129047660199805300971936 \\ 6 & -18.886678348734296934648840469575921 & -19.095091912250933709000748332754870 \\ 7 & -22.209056127812161704085770192533417 & -22.437790137971372572352358156488860 \\ 8 & -25.545482070626355796809539664033139 & -25.793196864919170855940747024418097 \end{tabular} } \end{table} \begin{table} \caption{\label{t3.6} Fit of the ansatz (\ref{3.F3}) to the data of Table \ref{t3.5} } \vspace{.4cm} {\small \begin{tabular}{l|l|l|lr|l|l|l|l} $N$ & $B_4^{\mathrm{hard}}$ & $K_4^{\mathrm{hard}}$ & $C_4^{\mathrm{hard}}$ & \quad & $B_4^{\mathrm{soft}}$ & $K_4^{\mathrm{hard}}$ & $C_4^{\mathrm{soft}}$ & $D_4^{\mathrm{soft}}$ \\ \hline 3,4,5,(6) & 0.749371& 0.059801& -0.091054 & & 0.497409& 0.120202& -0.052807&0.073687 \\ 4,5,6,(7) & 0.728988& 0.081365& -0.080181 && 0.509625& 0.099801& -0.040616& 0.040317 \\ 5,6,7,(8) & 0.723951& 0.087261& -0.078408 && 0.522124& 0.076905& -0.022675& -0.004884 \\ 6,7,8, (9) & 0.726340& 0.084219& -0.078810 && 0.521397& 0.078345& -0.024029 & -0.001556 \\ 7,8,9,(10) & 0.727263& 0.082957& -0.078795 && 0.518587& 0.084298& -0.030427 & 0.014202 \\ 8,9,10 & 0.726874& 0.083523& -0.078873 && & & & \\ \end{tabular} } \vspace{.4cm} {\small \begin{tabular}{l|l|l|lr|l|l|l|l} $N$ & $B_6^{\mathrm{hard}}$ & $K_6^{\mathrm{hard}}$ & $C_6^{\mathrm{hard}}$ & \qquad & $B_6^{\mathrm{soft}}$ & $K_6^{\mathrm{soft}}$ & $C_6^{\mathrm{soft}}$ \\ \hline 3,4,5 & 1.104506& -0.092158& -0.317518 && 0.967066& -0.059461& -0.248851 \\ 4,5,6 & 0.884919& 0.140146& -0.200388 && 0.795791& 0.121733& -0.157491 \\ 5,6,7 & 0.874984& 0.151776& -0.196890 && 0.786513& 0.132593& -0.154224 \\ 6,7,8 & 0.951461& 0.054407& -0.209757 && 0.842635& 0.061139& -0.163667 \\ \end{tabular} } \end{table} \subsection{Density and two-point distribution} \subsection*{Density} Consider for definiteness the disk geometry with a hard wall at $\Gamma = 4$. Using the formula (\ref{2.den}) the density profile can be calculated for up to 10 particles. One way to present the data is in graphical form with the boundary of the disk taken as the origin. This is done in Figure \ref{f3.1}. The plot shows rapid convergence of the profiles near the boundary. \begin{figure} \epsfbox{fig31.eps} \caption{\label{f3.1} Density profile in the hard disk case for several values of $N$ at $\Gamma=4$. The boundary of the disk is taken as origin. } \end{figure} To investigate the rate of convergence of the whole profile as measured from the boundary to the thermodynamic value we can investigate the contact theorem \cite{CFG80}. This expresses the thermodynamic pressure in terms of the density at contact with the wall, and the potential drop across the interface (which in turn is proportional to the first moment of the density profile). Explicitly the contact theorem states \begin{equation}\label{3.con} \Big ( 1 - {\Gamma \over 4} \Big ) \rho_b = \rho (0) - 2 \pi \rho_b \Gamma \int_0^\infty x (\rho (x) - \rho_b ) \, dx \end{equation} where we stress again that the density is measured from the boundary. Much to our initial surprise, the convergence of the r.h.s.~to the l.h.s.~for the finite $N$ data is very slow. For 10 particles the error is of order $30 \%$. Further investigation reveals that this is not special to the coupling $\Gamma = 4$. At $\Gamma = 2$ we have the analytic expression \cite{Ja81} $$ \rho_{(1)}(r) = {1 \over 2 \pi} \sum_{j=1}^N { (R - r)^{2j-2} e^{- \pi (R-r)^2} \over \int_0^R s^{2j - 1} e^{- \pi s^2} \, ds}, \quad 0 \le r \le R $$ where $r$ is measured from the boundary and the background density is taken to equal unity. Choosing $N=10$ and substituting in (\ref{3.con}) again gives an error of order $30\%$. Indeed choosing $N = 500$ still gives an error of order $3\%$. In fact the slow convergence of (\ref{3.con}) can be understood analytically by making use of a sum rule for the OCP applicable for the finite disk \cite{CFG80}. This sum rule reads \begin{equation}\label{3.sumr} \rho_{(1)}(0) - \Big ( 1 - {\Gamma \over 4} \Big ) \rho_b = - {\Gamma \rho_b^2 \pi^2 \over N} \int_0^R r^3 \Big (\rho_{(1)}(R - r) - \rho_b \Big ) \, dr \end{equation} where $\rho_{(1)}(r)$ is measured inward from the boundary. Noting that charge neutrality requires $$ \int_0^R r \Big ( \rho_{(1)}(R-r) - \rho_b \Big ) \, dr = \int_0^R ( R - r) \Big ( \rho_{(1)}(r) - \rho_b \Big ) \, dr $$ we can write \begin{eqnarray*}\lefteqn{ - {\Gamma \rho_b^2 \pi^2 \over N} \int_0^R r^3 \Big (\rho_{(1)}(R - r) - \rho_b \Big ) \, dr} \\ && = - {\Gamma \rho_b^2 \pi^2 \over N} \int_0^R (R - r)^3 \Big (\rho_{(1)}(r) - \rho_b \Big ) \, dr \\ && = - {\Gamma \rho_b^2 \pi^2 \over N} \int_0^R ( - 2r R^2 + 3r^2 R - r^3) \Big (\rho_{(1)}(r) - \rho_b \Big ) \, dr \\ && = 2 \Gamma \rho_b \pi \int_0^R r \Big (\rho_{(1)}(r) - \rho_b \Big ) \, dr - {3 \Gamma (\rho_b \pi)^{3/2} \over N^{1/2}} \int_0^R r^2 \Big (\rho_{(1)}(r) - \rho_b \Big ) \, dr \\ && \quad + {\Gamma \rho_b^2 \pi^2 \over N} \int_0^R r^3 \Big ( \rho_{(1)}(r) - \rho_b ) \, dr \end{eqnarray*} This shows that the finite size corrections to the r.h.s.~of (\ref{3.sumr}) are proportional to $N^{-1/2}$, thus explaining our empirical observation. \subsection*{Two-point function} At $\Gamma = 2$ and in the thermodynamic limit the two-particle distribution function has the exact evaluation \cite{Ja81} $$ \rho_{(2)}({0}, \vec{r}) = \rho_b^2 \Big ( 1 - e^{-\pi \rho_b |\vec{r}|^2} \Big ). $$ This is a monotonic function, with the corresponding truncated distribution $\rho_{(2)}^T({0}, \vec{r}) := \rho_{(2)}({0}, \vec{r}) - \rho_{(1)}({0})\rho_{(1)}(\r)$ exhibiting Gaussian decay to zero. There is evidence, both analytic and numerical \cite{Ja81,CLWH82} which suggests that for $\Gamma > 2$ the two-particle distribution exhibits oscillations. At $\Gamma = 4$ this feature has already been observed in the exact finite $N$ calculation of $ \rho_{(2)}({0}, \vec{r})$ by Samaj et al.~\cite{SPK94}. Furthermore, this feature should become more pronounced as $\Gamma$ increases. This is indeed what we observe when plotting our results for $\Gamma = 4$ and $\Gamma = 6$ on the same graph (see Figure \ref{f3.2}). \begin{figure} \epsfbox{fig32.eps} \caption{\label{f3.2} Two-point correlation in the sphere case for $N=8$ particles at $\Gamma=4$ and $\Gamma=6$. } \end{figure} The fact that the 2dOCP is a Coulomb system in its conductive phase implies that in the bulk the second moment of the truncated distribution obeys the Stillinger-Lovett sum rule \begin{equation}\label{3.still} \int_{{\mathbf{R}}^2} \vec{r}\,{}^2 \rho_{(2)}^T(0,\vec{r}) \, d\vec{r} = - {2 \over \pi \Gamma}. \end{equation} For the hard disk in the finite system we can compute \begin{equation}\label{3.stillf} \int_{|\vec{r}| < R} \vec{r}\,{}^2 \rho_{(2)}^T({0}, \vec{r}) \, d\vec{r} \end{equation} and compare it with the universal value given by (\ref{3.still}). At $\Gamma =4$ and with $N=9$ we find agreement with the universal value to within $2\%$. In fact, analogous to the integral in (\ref{3.sumr}), the integral (\ref{3.stillf}) can be evaluated exactly and the terms which differ from $-2/\pi \Gamma$ read off. In the hard wall case we find \begin{equation}\label{3.gs} \int_{|\vec{r}| < R} \r^2 \rho_{(2)}^T(\vec{0}, \vec{r}) \, d\vec{r} = - {2 \over \pi \Gamma} \bigg ( \rho_{(1)}(0) / \rho_b + N \rho_{(2)}^T(0,R)/ \rho_b^2 \bigg ), \end{equation} while in the soft wall case the same expression results except that the boundary term $N \rho_{(2)}^T(0,R)/ \rho_b^2$ is no longer present on the r.h.s., while on the l.h.s.~the integral is over ${\mathbf{R}}^2$. We see from (\ref{3.gs}) that the deviation in the finite system from the bulk value (\ref{3.still}) is determined by $$ - {2 \over \pi \Gamma} \Big ( (\rho_{(1)}(0) - \rho_b)/\rho_b + N \rho_{(2)}^T(0,R) / \rho_b^2 \Big ), $$ and thus consists of a bulk and surface contribution. \section{New sum rules}\label{sec:sum-rules} In this section we present the derivation of the sum rule~(\ref{3.gs}) and its generalization to multicomponent Coulomb systems. First we show that the sum rule can be derived within the formalism of Section~\ref{sec:formalism}, then we present a more general derivation of the sum rule. \subsection{The case $\Gamma$ even} The formalism presented in Section~\ref{sec:formalism} is valid only if $\Gamma$ is an even integer. Within this formalism we can use the expressions~(\ref{eq:2.rho2-g4}) and~(\ref{eq:2.rho2-g6}) for the two-point correlation functions (and its generalizations to higher $\Gamma$) to compute the second moment \begin{equation}\label{eq:4.sec-mom} \int_{\Lambda} \r^2 \rho_{(2)}^T(0,\r)\,d\r\,, \end{equation} where $\Lambda$ is a disk of radius $R$ (hard disk) or ${\mathbf{R}}^2$ (soft disk). For example in the hard disk case with $\Gamma=4$, for each term in the sum~(\ref{eq:2.rho2-g4}) the integral~(\ref{eq:4.sec-mom}) gives an incomplete gamma function $\gamma(\mu_j+2,2N)$. Then we use the recurrence relation \begin{equation}\label{eq:4.rec} \gamma(\mu_j+2,2N)=(\mu_j+1)\gamma(\mu_j+1,2N) -e^{-2N}(2N)^{\mu_j+1} \end{equation} to split the expression in two. The first term is proportional to $\rho_{(1)}(0)$ while the second is proportional to $\rho_{(2)}(0,R)$. The sum rule~(\ref{3.gs}) follows from that. The calculation can be easily generalized to any even $\Gamma$. In the soft disk case since the incomplete gamma functions are replaced by complete gamma functions the recurrence relation~(\ref{eq:4.rec}) does not have a second term on the r.h.s., therefore there is no surface contribution proportional to $\rho_{(2)}^T(0,R)$ in the sum rule. \subsection{General case} In fact a more general derivation of this sum rule, valid for any value of the coupling constant, can be obtained by studying the variations of the density as a function of the size of the disk. Let us consider the general case of a multicomponent jellium in $\nu$ dimensions confined in a spherical domain $\Lambda$ of radius $R$ and volume $V=\Omega_\nu R^\nu/\nu$ with $\Omega_\nu=2\pi^{\nu/2}/\Gamma(\nu/2)$. The system is composed of $s$ different species with charges $(e_\alpha)_{\alpha\in\{1,\dots,s\}}$ and there are $N_\alpha$ particles of the species $\alpha$. Let $N=\sum_\alpha N_\alpha$ be the total number of particles and let us define the average density of the species $\alpha$, $\rho_\alpha=N_\alpha/V$ and the total average density $\rho=N/V$. As in the preceding sections $\rho_b$ is the background number density and let $e_b$ be its charge so that the background charge density is $e_b\rho_b$. For convenience let us define the ``number of particles of the background'' by $N_b=\rho_b V$. In general the Coulomb potential is \begin{equation}\label{eq:defPhi} \Phi(\r)=\cases{-\ln r,&if $\nu=2$\cr \displaystyle\frac{r^{2-\nu}}{\nu-2},&otherwise,} \end{equation} and the Coulomb force is \begin{equation}\label{eq:defF} {\vec{F}}(\r)=-\nabla\Phi(\r)=\frac{\r}{r^\nu}\,. \end{equation} The Hamiltonian of the Coulomb system is \begin{equation}\label{eq:defU} U=\frac{1}{2}\sum_{i\neq j}e_{\alpha_i}e_{\alpha_j}\Phi(\r_i-\r_j) +e_b\rho_b\sum_{i=1}^N e_{\alpha_i}\int_\Lambda d\r \Phi(\r_i-\r) +\frac{e_b^2\rho_b^2}{2}\int_{\Lambda^2} d\r d\r' \Phi(\r-\r')\,. \end{equation} We shall consider the correlation functions in the canonical ensemble \begin{equation} \rho_{\alpha_1\dots\alpha_n}^{(n)}(\r_1,\dots,\r_n)=\left< \sum_{i_1=1}^{N_{\alpha_1}}\cdots \sum_{i_n=1}^{N_{\alpha_n}} \delta(\r_1-\r_{\alpha_1,i_1})\cdots \delta(\r_n-\r_{\alpha_n,i_n}) \right>\,. \end{equation} The $\left<\cdots\right>$ is the average in the canonical ensemble and in the preceding sums if some $\alpha_a=\alpha_b$ we exclude the term $i_a=i_b$ as usual. In three dimensions in order to have a well defined thermodynamic limit we shall restrict ourselves to the case where all electric charges $e_\alpha$ have the same sign and the background carries a opposite neutralizing charge. In two dimensions we can also consider systems with charges of different signs and eventually without background ($\rho_b=0$) if the coupling contants $|\beta e_\alpha e_\gamma| <2$ for all pair of charges $(e_\alpha,\, e_\gamma)$ of different signs. \subsubsection{Contact theorem sum rule} The derivation of the sum rule for the second moment of the two-point correlation function is similar to that of the contact theorem for a spherical domain~\cite{CFG80}. Let us first show here the generalization of this contact theorem for the multicomponent jellium. We consider the canonical partition function (times $N!$) \begin{equation} {\cal Q}=\int_{\Lambda^N} d\r^N \exp(-\beta U)\,, \end{equation} as a function of the volume $V$. We shall compute the thermodynamical pressure $p^{(\theta)}=\partial\log{\cal Q}/\partial V$ in two different ways. The derivative is done at fixed number of particles and fixed $N_b$. In general using the scaling $\r=V^{1/\nu}\tilde{\r}$ we have \begin{equation}\label{betap0} \beta p^{(\theta)}=\frac{\partial \ln {\cal Q}}{\partial V}= \rho-\frac{\beta V^N}{{\cal Q}}\int_{\tilde{\Lambda}^N} d\tilde{\r}^N \frac{\partial U(V^{1/\nu}\tilde{\r})}{\partial V} e^{-\beta U}\,, \end{equation} where $\tilde{\Lambda}$ is a sphere of volume 1. A first way to compute the derivative of $U$ is by using the general formula \begin{equation}\label{eq:virial} \frac{\partial \Phi(V^{1/\nu}\tilde{\r})}{\partial V}= -\frac{1}{\nu V}\,\r\cdot {\vec{F}}(\r)\,. \end{equation} This gives, together with the definition~(\ref{eq:defU}) of $U$, \begin{eqnarray}\label{eq:betap1} \beta p^{(\theta)}&=\rho+ \displaystyle\frac{\beta}{\nu V}& \left[ \int_{\Lambda^2}d\r d\r' \r\cdot {\vec{F}}(\r-\r')\sum_{\alpha,\alpha'} e_\alpha e_{\alpha'}\rho_{\alpha\alpha'}^{(2)}(\r,\r') \right. \nonumber\\ && +e_b\rho_b\int_{\Lambda^2}d\r d\r' (\r-\r')\cdot {\vec{F}}(\r-\r') \sum_\alpha e_\alpha \rho_\alpha^{(1)}(\r') \nonumber\\ && \left. +\frac{1}{2}e_b^2\rho_b^2 \int_{\Lambda^2} d\r d\r' (\r-\r')\cdot {\vec{F}}(\r-\r') \right]\,. \end{eqnarray} We can transform the preceding expression by using the first equation of the BGY hierarchy \begin{equation}\label{eq:BGY1} k_B T\nabla \rho_\alpha(\r)=e_\alpha \rho_b e_b\int_\Lambda d\r'\, {\vec{F}}(\r-\r')\rho_\alpha^{(1)}(\r)+ \int_\Lambda d\r' \sum_{\alpha'} e_\alpha e_{\alpha'} {\vec{F}}(\r-\r') \rho_{\alpha \alpha'}^{(2)} (\r,\r)\,. \end{equation} The r.h.s of~(\ref{eq:BGY1}) appears in the first and second lines of~(\ref{eq:betap1}). Replacing it by the l.h.s of~(\ref{eq:BGY1}) we find \begin{eqnarray} \beta p^{(\theta)}&=\rho+ \displaystyle\frac{\beta}{\nu V}& \left[ k_B T \int_{\Lambda} d\r \sum_\alpha \r\cdot\nabla\rho_\alpha^{(1)}(\r) \right. \nonumber\\ && -e_b \rho_b \int_{\Lambda^2}d\r d\r' \r'\cdot{\vec{F}}(\r-\r') \sum_\alpha e_\alpha \rho_\alpha^{(1)}(\r) \nonumber\\ && \left. +\frac{1}{2} \rho_b^2 e_b^2 \int_{\Lambda^2} d\r d\r'\, (\r-\r')\cdot {\vec{F}}(\r-\r') \right]\,. \end{eqnarray} The first term of the r.h.s of the preceding equation can be computed by integration by parts while the others can be computed using the definition~(\ref{eq:defF}) of the Coulomb force ${\vec{F}}$ and Newton's theorem. This yields the following expression for the thermodynamical pressure \begin{eqnarray}\label{eq:betapfin1} \beta p^{(\theta)}&=&\sum_\alpha \rho_\alpha^{(1)}(R) +\beta e_b \rho_b R^{2-\nu}\left(\sum_\alpha e_\alpha \frac{N_\alpha}{2} +e_b\frac{N_b}{\nu+2}\right) \nonumber\\ && -\frac{\beta\rho_b e_b}{2R^\nu} \int_\Lambda d\r\, r^2 \sum_\alpha e_\alpha \rho_\alpha^{(1)}(\r) \,. \end{eqnarray} The other way to compute the thermodynamical pressure is to use the actual scaling properties of the Coulomb potential $\Phi$, \begin{equation}\label{eq:scalPhi} \frac{\partial\Phi(V^{1/\nu}\tilde{\r})}{\partial V} =\cases{\displaystyle-\frac{1}{2V},&if $\nu=2$\cr \displaystyle\frac{2-\nu}{\nu V}\,\Phi(\r),&otherwise. } \end{equation} Substituting this expression in~(\ref{betap0}) gives \begin{equation}\label{eq:betapfin2} \beta p^{(\theta)}=\rho+ {\beta \delta_{\nu,2} \over 4} \left(\frac{ Q^2}{V} -\sum_\alpha e_\alpha^2 \rho_\alpha \right) + \frac{\nu-2}{\nu V}\beta \left<U\right>\,, \end{equation} where $Q=\sum_\alpha e_\alpha N_\alpha+ e_b N_b$ is the total charge of the system. Equating the two expressions~(\ref{eq:betapfin1}) and~(\ref{eq:betapfin2}) of the thermodynamic pressure we find the generalization of the contact theorem \begin{eqnarray}\label{eq:contact} \lefteqn{ \sum_\alpha \rho_\alpha^{(1)}(R)+\frac{\beta e_b \rho_b}{2R^{\nu-2}}Q -\frac{\beta e_b \rho_b}{2R^\nu}\int_\Lambda d\r\, r^2 q(\r) } \nonumber\\ &&= \rho + \delta_{\nu,2} \frac{\beta}{4} \left(\frac{Q^2}{V}-\sum_\alpha e_\alpha^2 \rho_\alpha\right) +\frac{\nu-2}{\nu V} \beta \left<U\right> \end{eqnarray} where $q(\r)=\sum_\alpha e_\alpha \rho_\alpha^{(1)}(\r)+e_b \rho_b$ is the local charge density. \subsubsection{Density-charge correlation second moment sum rule} Similar calculations lead to the second moment sum rule for the density-charge truncated correlation function $\sum_\beta e_\beta\rho_{\alpha\beta}^{(2)T}(0,\r)$. Here we consider the quantity \begin{equation} {\cal Q}_\alpha = \int_{\Lambda^N} d\r^N e^{-\beta U} \sum_{i=1}^{N_\alpha} \delta(\r_{i,\alpha}) \,, \end{equation} as a function of the volume $V$. Note that the density of the species $\alpha$ at the center of the spherical domain is $\rho_\alpha^{(1)}(0)={\cal Q}_\alpha/{\cal Q}$. Like in the preceding section we want to compute by two different ways the quantity ${\cal Q}^{-1}\partial {\cal Q}_\alpha/\partial V$. Using the same scaling argument as before we have \begin{equation}\label{eq:dQalpha0} \frac{1}{{\cal Q}}\frac{\partial{\cal Q}_\alpha}{\partial V}= \frac{N-1}{V}\rho_\alpha^{(1)}(0) -\beta \frac{V^N}{{\cal Q}} \int_{\tilde{\Lambda}^N} d\tilde{r}^N \sum_{i=1}^{N_\alpha} \delta(V^{1/\nu}\tilde{\r}_{i,\alpha}) \,\frac{\partial U(V^{1/\nu}\tilde{\r})}{\partial V} \,e^{-\beta U} \,. \end{equation} Using eq.~(\ref{eq:virial}) and the definition~(\ref{eq:defU}) of the Hamiltonian $U$ we find \begin{eqnarray}\label{eq:qalpha1} \frac{1}{{\cal Q}}\frac{\partial{\cal Q}_\alpha}{\partial V}&=& \frac{N-1}{V}\rho_\alpha^{(1)}(0) \nonumber\\ &&+\frac{\beta}{\nu V} \left[ \int_{\Lambda^2}d\r d\r'\, \r\cdot{\vec{F}}(\r-\r') \sum_{\beta,\gamma} e_\beta e_\gamma \rho_{\alpha\beta\gamma}^{(3)} (0,\r,\r') \right. \nonumber\\ &&+\int_\Lambda \r\cdot{\vec{F}}(\r) e_\alpha \sum_\beta e_\beta \rho_{\alpha\beta}^{(2)}(0,\r) \nonumber\\ &&+e_b \rho_b \int_{\Lambda^2}d\r d\r'\sum_\beta e_\beta \rho_{\alpha\beta}^{(2)}(0,\r)\, \r\cdot{\vec{F}}(\r-\r') \nonumber\\ &&-e_b \rho_b \int_{\Lambda^2} d\r d\r' \r'\cdot{\vec{F}}(\r-\r') \sum_\beta e_\beta \rho_{\alpha\beta}^{(2)}(0,\r) \nonumber\\ &&+e_b \rho_b\int_\Lambda d\r\,\r\cdot{\vec{F}}(\r) e_\alpha \rho_\alpha^{(1)}(0) \nonumber\\ && \left. +\frac{1}{2} e_b^2\rho_b^2\int_{\Lambda^2} d\r d\r'\, (\r-\r')\cdot{\vec{F}}(\r-\r') \rho_\alpha^{(1)}(0) \right] \end{eqnarray} Using the second BGY equation \begin{eqnarray}\label{eq:BGY2} k_B T\nabla_\r\rho_{\alpha\beta}^{(2)}(0,\r)&=&e_\beta e_b \rho_b \int_\Lambda d\r'\,{\vec{F}}(\r-\r')\rho_{\alpha\beta}^{(2)}(0,\r) \nonumber\\ &&+e_\beta e_\alpha {\vec{F}}(\r) \rho_{\alpha\beta}^{(2)}(0,\r) \nonumber\\ &&+\int_\Lambda d\r'\,{\vec{F}}(\r-\r')\sum_\gamma e_\beta e_\gamma \rho_{\alpha\beta\gamma}^{(3)}(0,\r,\r') \,, \end{eqnarray} and then integration by parts \begin{equation} \sum_\beta \int_\Lambda \r\cdot\nabla_\r\rho_{\alpha\beta}^{(2)} (0,\r)=\nu V\sum_\beta \rho_{\alpha\beta}^{(2)}(0,R) -\nu(N-1)\rho_\alpha^{(1)}(0) \,, \end{equation} we can arrange expression~(\ref{eq:qalpha1}) to find, after computing explicitly the integrals involving ${\vec{F}}$ using Newton's theorem, \begin{eqnarray}\label{eq:dQalpha1} \frac{1}{{\cal Q}}\frac{\partial{\cal Q}_\alpha}{\partial V}&=& \sum_\beta \rho_{\alpha\beta}^{(2)}(0,R) -\frac{\beta e_b\rho_b}{2R^\nu} \int_\Lambda d\r\,r^2\sum_\beta e_\beta \rho_{\alpha\beta}^{(2)}(0,\r) \nonumber\\ &&+\frac{\beta e_b\rho_b}{2R^{\nu-2}}\rho_\alpha^{(1)}(0) \left[\sum_\beta e_\beta N_\beta+\frac{e_b N_b}{\nu+2}\right] \,. \end{eqnarray} The second way for computing ${\cal Q}^{-1}\partial{\cal Q}_\alpha/\partial V$ is by using directly equation~(\ref{eq:scalPhi}) into equation~(\ref{eq:dQalpha0}). This gives, \begin{eqnarray}\label{eq:dQalpha2} \frac{1}{{\cal Q}}\frac{\partial{\cal Q}_\alpha}{\partial V}&=& \frac{N-1}{V}\rho_\alpha^{(1)}(0)+\frac{\beta}{4}\delta_{\nu,2} \left[\frac{Q^2}{V}-\sum_\beta e_\beta^2 N_\beta\right]\rho_\alpha^{(1)}(0) \nonumber\\ && +\frac{\nu-2}{\nu V}\beta \left< U\hat{\rho}_\alpha^{(1)}(0) \right> \,, \end{eqnarray} where $\hat{\rho}_\alpha^{(1)}(0)=\sum_{i=1}^{N_\alpha}\delta(\r_{i,\alpha})$ is the microscopic density of $\alpha$-particles at the center of the domain $\Lambda$. Comparing the two expressions~(\ref{eq:dQalpha1}) and~(\ref{eq:dQalpha2}) of ${\cal Q}^{-1}\partial{\cal Q}_\alpha/\partial V$ gives a sum rule for the second moment of the density of $\alpha$ particles-electric charge correlation function. The sum rule takes a nice form by considering the truncated correlation function and making use of the contact sum rule~(\ref{eq:contact}), \begin{eqnarray} \label{eq:thenewsumrule} \frac{\beta e_b \rho_b\Omega_\nu}{2\nu} \int_{\Lambda} d\r\, r^2 \sum_\beta e_\beta \rho_{\alpha\beta}^{(2)T}(0,\r) &=& \rho_\alpha^{(1)}(0) +\frac{\Omega_\nu}{\nu} R^\nu \sum_\beta \rho_{\alpha\beta}^{(2)T}(0,R) \nonumber\\ &&+\frac{2-\nu}{\nu}\, \beta\left<U\hat{\rho}_{\alpha}^{(1)}(0)\right>^T \,. \end{eqnarray} In the case of the two-dimensional OCP ($\nu=2$, $s=1$ and $e_b=-e_1$) this is exactly the sum rule~(\ref{3.gs}) announced in the preceding section $$ \int_{|\vec{r}| < R} \r^2 \rho_{(2)}^T(\vec{0}, \vec{r}) \, d\vec{r} = - {2 \over \pi \Gamma} \bigg ( \rho_{(1)}(0) / \rho_b + N \rho_{(2)}^T(0,R)/ \rho_b^2 \bigg ) \,. \eqno\hbox{(\ref{3.gs})} $$ The sum rule~(\ref{eq:thenewsumrule}) is in fact a series of $s$ sum rules for the density-charge correlation function $\sum_\beta e_\beta \rho^{(2)T}_{\alpha\beta}(0,\r)$ for each species $\alpha$. By taking the sum of these sum rules with the factors $e_\alpha$, we find a sum rule for the charge-charge truncated correlation function $S(0,\r)=\sum_{\alpha,\beta} e_\alpha e_\beta \rho^{(2)T}_{\alpha\beta}(0,\r)$, \begin{eqnarray} \label{eq:charge-sumrule} \frac{\beta e_b \rho_b\Omega_\nu}{2\nu} \int_{\Lambda} d\r\, r^2 S(0,\r) &=& \sum_{\alpha}e_\alpha\rho_\alpha^{(1)}(0) +\frac{\Omega_\nu}{\nu} R^\nu \sum_{\alpha,\beta} e_\alpha \rho_{\alpha\beta}^{(2)T}(0,R) \nonumber\\ &&+\frac{2-\nu}{\nu}\, \beta\left<U\sum_\alpha e_\alpha \hat{\rho}_{\alpha}^{(1)}(0)\right>^T \,. \end{eqnarray} \subsubsection{Thermodynamic limit of the sum rules} \subsubsection*{Canonical ensemble} In order to study the relationship between sum rules~(\ref{eq:thenewsumrule}) and~(\ref{eq:charge-sumrule}) and the Stillinger--Lovett sum rule, we need to know the behavior of the correlation functions as they approach the thermodynamic limit. This behavior is different depending on the ensemble used. In this section we continue to work in the canonical ensemble. In general we shall suppose that in the thermodynamic limit the system is in a fluid and conducting phase. In this case the density becomes uniform in the thermodynamic limit $\rho_\alpha^{(1)}(0)\to \rho_\alpha$ and \begin{equation}\label{eq:Urho} \left<\hat{\rho}_\alpha^{(1)}(0)U\right>^T \to\left<\rho_\alpha U\right>^T=0\,, \end{equation} because in the canonical ensemble the density does not fluctuate. Let us first consider the case of a multicomponent Coulomb system without background (in two dimensions with small Coulomb couplings). In that case equation~(\ref{eq:thenewsumrule}) becomes \begin{equation}\label{eq:can-rhob-0-fini} \rho_\alpha^{(1)}(0) +\frac{\Omega_\nu}{\nu} R^\nu \sum_\beta \rho_{\alpha\beta}^{(2)T}(0,R) =0\,. \end{equation} This equation~(\ref{eq:can-rhob-0-fini}) give us the behavior of the correlation functions as they approach the thermodynamic limit \begin{equation}\label{eq:tails} \sum_{\gamma} \rho_{\alpha\gamma}^{(2)T} (0,R) \sim -\rho\rho_\alpha/N \,. \end{equation} This is a generalization of an already known result concerning the existence of $1/N$ tails for the correlation functions of one component fluids with short range forces~\cite{LP61}. However, for a neutral system taking the sum of equations~(\ref{eq:tails}) with the coefficients $e_\alpha$ show that the charge-total density correlation does not have $1/N$ tails, \begin{equation} R^\nu\sum_{\alpha\gamma} e_\alpha \rho_{\alpha\gamma}^{(2)T} (0,R) \to 0 \,, \end{equation} It is likely that a similar behavior exists in the general case ($\rho_b\neq0$, $\nu=2,3$), so it would be difficult to derive from~(\ref{eq:thenewsumrule}) partial sum rules for the density-charge correlations in the thermodynamic limit because with the $1/N$ tails, one cannot commute the thermodynamic limit with the integration over the space. However, one can conjecture that although the density-density correlations have $1/N$ tails, in the conductive phase the total density-charge correlations do not have these tails as it is in the case when $\rho_b=0$. If this is true, and assuming that the convergence of the charge-charge correlation function is uniform (in order to commute the thermodynamic limit with the integration over the space), one can recover the Stillinger--Lovett sum rule from the sum rule~(\ref{eq:charge-sumrule}) for finite systems, \begin{equation}\label{eq:SL} \frac{\beta\Omega_\nu}{2\nu} \int_{{\mathbf{R}}^\nu} d\r r^2 S(0,\r)=-1\,. \end{equation} The fact that we recover the Stillinger--Lovett sum rule is of course not a proof of our conjecture, but at least it show that our conjecture is not in contradiction with well known results. \subsubsection*{Grand canonical ensemble} For systems with short range forces the correlations functions do not have $1/N$ tails in the grand canonical ensemble as they approach the thermodynamic limit~\cite{LP61}. We will show that this is also the case for two-dimensional Coulomb systems with small couplings when there is no charged background and assuming this is also the case in general for a multicomponent jellium we will discuss the thermodynamic limit of the partial sum rules. The partial sum rules~(\ref{eq:thenewsumrule}) obtained before are different in the grand canonical ensemble. The grand canonical ensemble is parametrized by the background density $\rho_b$ and $s-1$ fugacities $\{z_\gamma\}$ used to fix $s-1$ average densities $\rho_\gamma$, the remaining density fixed by electroneutrality. The grand canonical version of the sum rules~(\ref{eq:thenewsumrule}) can be obtained in a straightforward manner by adapting the calculations of the last section. However special care should be taken because of the fluctuation of the average densities in the grand canonical ensemble. These fluctuations add some extra terms to sum rule~(\ref{eq:thenewsumrule}), \begin{eqnarray} \label{eq:sumrule-gc} \frac{\beta e_b \rho_b\Omega_\nu}{2\nu} \int_{\Lambda} d\r\, r^2 \sum_\beta e_\beta \rho_{\alpha\beta}^{(2)T}(0,\r) &=& \left<\hat\rho_\alpha^{(1)}(0)\right> +\frac{\Omega_\nu}{\nu} R^\nu \sum_\beta \rho_{\alpha\beta}^{(2)T}(0,R) \nonumber\\ &&-\left<N\hat\rho^{(1)}_\alpha(0)\right>^T +\frac{2-\nu}{\nu}\, \beta\left<U\hat{\rho}_{\alpha}^{(1)}(0)\right>^T \nonumber\\ &&+\delta_{\nu,2} \left<\sum_\gamma \frac{\beta e_\gamma^2}{4} N_\gamma \hat\rho_\alpha^{(1)}(0)\right>^T \,. \end{eqnarray} To proceed with the discussion of the thermodynamic limit of this sum rule, we need to use a relation that will allow us to simplify the terms on the r.h.s.~of equation~(\ref{eq:sumrule-gc}) in the thermodynamic limit. This relation reads for $\nu=2$ or 3, \begin{equation}\label{eq:scaling-relation} \rho_\alpha -\left<N\rho_\alpha\right>^T +\frac{2-\nu}{\nu}\, \beta\left<U\rho_{\alpha}\right>^T +\delta_{\nu,2} \left<\sum_\gamma \frac{\beta e_\gamma^2}{4} N_\gamma \rho_\alpha\right>^T = \rho_b\frac{\partial\rho_\alpha}{\partial\rho_b} \,. \end{equation} This relation is a consequence of the scaling properties of the Coulomb potential. To prove it, let us consider the thermodynamic grand canonical pressure \begin{equation} \beta \tilde p(\beta,\{z_\gamma\},\rho_b) =\lim_{V\to\infty} V^{-1} \ln \Xi(\beta,\{z_\gamma\},\rho_b,V) \,, \end{equation} where $\Xi$ is the grand canonical partition function. Using the scaling properties of the Coulomb potential we have for $\nu=3$ \begin{equation} \beta \tilde p(\beta,\{z_\alpha\},\rho_b)=\lambda^4\beta \tilde p(\lambda\beta,\{\lambda^{-3/2}z_\alpha\},\lambda^{-3}\rho_b) \,, \end{equation} and for $\nu=2$, \begin{equation} \beta \tilde p(\beta,\{z_\alpha\},\rho_b)= \beta \tilde p(\beta,\{\lambda^{-2(1-(\beta e_\alpha^2/4))}z_\alpha\}, \lambda^{-2}\rho_b) \,, \end{equation} for any positive number $\lambda$. Taking the derivative of these relations with respect to $\lambda$, then putting $\lambda=1$ and using the usual thermodynamic relations yields for $\nu=3$, \begin{equation}\label{eq:p-nu3} \tilde p=\frac{1}{3}\left<H\right> + \frac{1}{2\beta} \rho - \rho_b\frac{\partial\tilde p}{\partial\rho_b} =\frac{1}{3}\left<U\right>+ \frac{1}{\beta} \rho -\rho_b\frac{\partial\tilde p}{\partial\rho_b} \,, \end{equation} and for $\nu=2$, \begin{equation}\label{eq:p-nu2} \beta \tilde p=\sum_\alpha \left(1-\frac{\beta e_\alpha^2}{4}\right) \rho_\alpha +\beta\rho_b \frac{\partial\tilde p}{\partial\rho_b}\,. \end{equation} where $\left<H\right>$ is the total internal energy (including the kinetic term). The announced relation~(\ref{eq:scaling-relation}) follows from taking the derivative of~(\ref{eq:p-nu3}) and~(\ref{eq:p-nu2}) with respect to the fugacities. As before let us consider first the case $\rho_b=0$ (in two dimensions for systems with small couplings). Then equation~(\ref{eq:sumrule-gc}) together with equation~(\ref{eq:scaling-relation}) shows that the grand canonical total density-partial density correlation function does not exhibit any $1/N$ tails, \begin{equation}\label{eq:notails} R^\nu \sum_{\alpha\gamma} \rho^{(2)T}_{\alpha\gamma}(0,R) \to0\,. \end{equation} Now if we suppose that in the general case ($\rho_b\neq0$, $\nu=2,3$) this property still holds and that the density-charge correlation functions converge uniformly we recover the partial sum rules \begin{equation}\label{eq:sumrule-svw} \frac{\beta e_b\Omega_\nu}{2\nu} \int_{{\mathbf{R}}^\nu} d\r r^2 \sum_\gamma e_\gamma\rho_{\alpha\gamma}^{(2)T} (0,\r) = \frac{\partial\rho_\alpha}{\partial\rho_b} \,, \end{equation} that have been previously derived by Suttorp and van~Wonderen~\cite{SvW87} in the three dimensional case. These equations also hold for two-dimensional systems. One can recover the Stillinger--Lovett sum rule~(\ref{eq:SL}) by taking the sum of these equations~(\ref{eq:sumrule-svw}) with the factors $e_\alpha$ and using electroneutrality. Notice that the condition~(\ref{eq:notails}) on the thermodynamic limit of the two-point correlation function when one of the points is in the boundary is different from the usual condition needed to prove the Stillinger--Lovett~\cite{MG83} that the correlation function of the infinite system should decay faster than $1/r^{\nu+2}$. Notwithstanding the relation of the sum rules~(\ref{eq:thenewsumrule}) and~(\ref{eq:sumrule-gc}) with the Stillinger--Lovett sum rule~(\ref{eq:SL}), let us stress that for finite systems these sum rules are not screening sum rules like the Stillinger--Lovett sum rule since for finite systems the screening of external charges does not exists (because since the total electric charge is conserved, the excess of charge can not leak out to infinity like it does in infinite systems). From the derivation presented in the previous section it is clear that the new sum rules should be seen more as a second order contact theorem rather than a screening sum rule. Futhermore when there is no background ($\rho_b=0$) the relation with Stillinger--Lovett sum rule disappears because the term containing the second moment of the density-charge correlation vanishes. \section{Summary and conclusion} Expanding the power of the Vandermonde determinant that appears in the Boltzmann factor of the 2dOCP in terms of simple orthogonal polynomials we have been able to develop exact numerical solutions for values of the coupling constant $\Gamma=4$ and $\Gamma=6$ for finite systems up to 11 and 9 particles respectively for different kinds of geometry (sphere, soft and hard wall disk). With these solutions we have been able to test the prediction~\cite{JMP94} of universal logarithmic finite size corrections to the free energy~(\ref{1.F}). Studying the correlation functions has lead us to find a new sum rule~(\ref{3.gs}) similar to the Stillinger--Lovett sum rule for finite systems. This sum rule can be derived within the formalism of section~\ref{sec:formalism}, but can also be generalized to higher dimension and multicomponent jellium systems (eq.~(\ref{eq:thenewsumrule})). Further applications of the formalism presented here are the study of surface correlations which are expected to have a universal behavior at large distances~\cite{Jan95}. Also the formal expressions of the correlations functions~(\ref{eq:2.rho2-g4}) and~(\ref{eq:2.rho2-g6}) could eventually be used to find higher order sum rules or other general properties. \section*{Acknowledgements} G.~T.~acknowledges the financial support from the Australian Research Council and would like to thank the Department of Mathematics and Statistics of the University of Melbourne for its hospitality. Also, we are particularly grateful to J.-Y.~Thibon and B.G.~Wybourne for supplying us with their data from \cite{STW94}. We thank B.~Jancovici for a useful remark on section~\ref{sec:sum-rules}.
\section{\@startsection{section}{1}{\zeta@}{3.5ex plus 1ex minus .2ex}{2.3ex plus .2ex}{\large\bf}} \def\arabic{section}.{\arabic{section}.} \def\Roman{section}-\Alph{subsection}.{\Roman{section}-\Alph{subsection}.} \def#1}{} \def\FERMILABPub#1{\def#1}{#1}} \def\ps@headings{\def\@oddfoot{}\def\@evenfoot{} \def\@oddhead{\hbox{}\hfill \makebox[.5\textwidth]{\raggedright\ignorespaces --\thepage{}-- \hfill }} \def\@evenhead{\@oddhead} \def\subsectionmark##1{\markboth{##1}{}} } \ps@headings \catcode`\@=12 \relax \def\rho#1{\ignorespaces $^{#1}$}{ \def\figcap{\section*{Figure Captions\markboth {FIGURECAPTIONS}{FIGURECAPTIONS}}\list {Fig. \arabic{enumi}:\hfill}{\settowidth\labelwidth{Fig. 999:} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endfigcap\endlist \relax \def\tablecap{\section*{Table Captions\markboth {TABLECAPTIONS}{TABLECAPTIONS}}\list {Table \arabic{enumi}:\hfill}{\settowidth\labelwidth{Table 999:} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endtablecap\endlist \relax \def\reflist{\section*{References\markboth {REFLIST}{REFLIST}}\list {[\arabic{enumi}]\hfill}{\settowidth\labelwidth{[999]} \leftmargin\labelwidth \advance\leftmargin\labelsep\usecounter{enumi}}} \let\endreflist\endlist \relax \catcode`\@=11 \def\marginnote#1{} \newcount\hour \newcount\minute \newtoks\amorpm \hour=\time\divide\hour by60 \minute=\time{\multiply\hour by60 \global\advance\minute by- \hour} \edef\standardtime{{\ifnum\hour<12 \global\amorpm={am}% \else\global\amorpm={pm}\advance\hour by-12 \fi \ifnum\hour=0 \hour=12 \fi \number\hour:\ifnum\minute<100\fi\number\minute\the\amorpm}} \edef\militarytime{\number\hour:\ifnum\minute<100\fi\number\minute} \def\NPB#1#2#3{Nucl. Phys. { B#1} (19#2) #3} \def\PLB#1#2#3{Phys. Lett. { B#1} (19#2) #3} \def\PLBold#1#2#3{Phys. Lett. {#1B} (19#2) #3} \def\PRD#1#2#3{Phys. Rev. { C#1} (19#2) #3} \def\PRL#1#2#3{Phys. Rev. Lett. {#1} (19#2) #3} \def\PRT#1#2#3{Phys. Rep. { C#1} (19#2) #3} \def\ARAA#1#2#3{Ann. Rev. Astron. Astrophys. {#1} (19#2) #3} \def\ARNP#1#2#3{Ann. Rev. Nucl. Part. Sci. {#1} (19#2) #3} \def\MODA#1#2#3{Mod. Phys. Lett. { A#1} (19#2) #3} \def\draftlabel#1{{\@bsphack\if@filesw {\let\thepage\relax \xdef\@gtempa{\write\@auxout{\string \newlabel{#1}{{\@currentlabel}{\thepage}}}}}\@gtempa \if@nobreak \ifvmode\nobreak\fi\fi\fi\@esphack} \gdef\@eqnlabel{#1}} \def\@eqnlabel{} \def\@vacuum{} \def\draftmarginnote#1{\marginpar{\raggedright\scriptsize\tt#1}} \def\draft{\oddsidemargin -.5truein \def\@oddfoot{\sl preliminary draft \hfil \rm\thepage\hfil\sl\today\quad\militarytime} \let\@evenfoot\@oddfoot \overfullrule 3pt \let\label=\draftlabel \let\marginnote=\draftmarginnote \def\@eqnnum{(\theequation)\rlap{\kern\marginparsep\tt\@eqnlabel}% \global\let\@eqnlabel\@vacuum} } \def\preprint{\twocolumn\sloppy\flushbottom\parindent 1em \leftmargini 2em\leftmarginv .5em\leftmarginvi .5em \oddsidemargin -.5in \evensidemargin -.5in \columnsep 15mm \footheight 0pt \textwidth 250mmin \topmargin -.4in \headheight 12pt \topskip .4in \textheight 175mm \footskip 0pt \def\@oddhead{\thepage\hfil\addtocounter{page}{1}\thepage} \let\@evenhead\@oddhead \def\@oddfoot{} \def\@evenfoot{} } \def\titlepage{\@restonecolfalse\if@twocolumn\@restonecoltrue\onecolumn \else \newpage \fi \thispagestyle{empty}\chi@page\zeta@ \def\arabic{footnote}{\fnsymbol{footnote}} } \def\endtitlepage{\if@restonecol\twocolumn \else \fi \def\arabic{footnote}{\arabic{footnote}} \setcounter{footnote}{0}} \catcode`@=12 \relax \def#1}{} \def\FERMILABPub#1{\def#1}{#1}} \def\ps@headings{\def\@oddfoot{}\def\@evenfoot{} \def\@oddhead{\hbox{}\hfill \makebox[.5\textwidth]{\raggedright\ignorespaces --\thepage{}-- \hfill }} \def\@evenhead{\@oddhead} \def\subsectionmark##1{\markboth{##1}{}} } \ps@headings \relax \newcommand{\ampl}[2]{{\cal M}\left( #1 \to #2 \right)} \newcommand{\mbox{d}}{\mbox{d}} \let\under=\beta \let\ced=\chi \let\du=\delta \let\um=\H \let\sll=\lambda \let\Sll=\Lambda \let\slo=\omega \let\Slo=\Omega \let\tie=\tau \let\br=\upsilon \def\alpha{\alpha} \def\beta{\beta} \def\chi{\chi} \def\delta{\delta} \def\epsilon{\epsilon} \def\frac{\phi} \def\gamma{\gamma} \def\eta{\eta} \def\iota{\iota} \def\psi{\psi} \def\kappa{\kappa} \def\lambda{\lambda} \def\mu{\mu} \def\nu{\nu} \def\omega{\omega} \def\pi{\pi} \def\theta{\theta} \def\rho{\rho} \def\sigma{\sigma} \def\tau{\tau} \def\upsilon{\upsilon} \def\xi{\xi} \def\zeta{\zeta} \def\Delta{\Delta} \def\Phi{\Phi} \def\Gamma{\Gamma} \def\Psi{\Psi} \def\Lambda{\Lambda} \def\Omega{\Omega} \def\Pi{\Pi} \def\Theta{\Theta} \def\Sigma{\Sigma} \def\Upsilon{\Upsilon} \def\Xi{\Xi} \def\varphi{\varphi} \def\varrho{\varrho} \def\varsigma{\varsigma} \def\vartheta{\vartheta} \def{\cal A}{{\cal A}} \def{\cal B}{{\cal B}} \def{\cal C}{{\cal C}} \def{\cal D}{{\cal D}} \def{\cal E}{{\cal E}} \def{\cal F}{{\cal F}} \def{\cal G}{{\cal G}} \def{\cal H}{{\cal H}} \def{\cal I}{{\cal I}} \def{\cal J}{{\cal J}} \def{\cal K}{{\cal K}} \def{\cal L}{{\cal L}} \def{\cal M}{{\cal M}} \def{\cal N}{{\cal N}} \def{\cal O}{{\cal O}} \def{\cal P}{{\cal P}} \def{\cal Q}{{\cal Q}} \def{\cal R}{{\cal R}} \def{\cal S}{{\cal S}} \def{\cal T}{{\cal T}} \def{\cal U}{{\cal U}} \def{\cal V}{{\cal V}} \def{\cal W}{{\cal W}} \def{\cal X}{{\cal X}} \def{\cal Y}{{\cal Y}} \def{\cal Z}{{\cal Z}} \def\Sc#1{{\hbox{\sc #1}}} \def\Sf#1{{\hbox{\sf #1}}} \def\slpa{\slash{\pa}} \def\slin{\SLLash{\in}} \def\bo{{\raise.15ex\hbox{\large$\Box$}}} \def\cbo{\Sc [} \def\pa{\partial} \def\de{\nabla} \def\dell{\bigtriangledown} \def\su{\sum} \def\pr{\prod} \def\iff{\leftrightarrow} \def\conj{{\hbox{\large *}}} \def\ \raisebox{-.4ex}{\rlap{$\sim$}} \raisebox{.4ex}{$<$}\ {\raisebox{-.4ex}{\rlap{$\sim$}} \raisebox{.4ex}{$<$}} \def\gtap{\raisebox{-.4ex}{\rlap{$\sim$}} \raisebox{.4ex}{$>$}} \def\face{{\raise.2ex\hbox{$\displaystyle \bigodot$}\mskip-2.2mu \llap {$\ddot \smile$}}} \def\ddg{\ddagger} \def\sp#1{{}^{#1}} \def\sb#1{{}_{#1}} \def\oldsl#1{\rlap/#1} \def\slash#1{\rlap{\hbox{$\mskip 1 mu /$}}#1} \def\Slash#1{\rlap{\hbox{$\mskip 3 mu /$}}#1} \def\SLash#1{\rlap{\hbox{$\mskip 4.5 mu /$}}#1} \def\SLLash#1{\rlap{\hbox{$\mskip 6 mu /$}}#1} \def\Tilde#1{\widetilde{#1}} \def\Hat#1{\widehat{#1}} \def\Bar#1{\overline{#1}} \def\bra#1{\left\langle #1\right|} \def\ket#1{\left| #1\right\rangle} \def\VEV#1{\left\langle #1\right\rangle} \def\abs#1{\left| #1\right|} \def\leftrightarrowfill{$\mathsurround=0pt \mathord\leftarrow \mkern-6mu \cleaders\hbox{$\mkern-2mu \mathord- \mkern-2mu$}\hfill \mkern-6mu \mathord\rightarrow$} \def\dvec#1{\vbox{\ialign{##\crcr \leftrightarrowfill\crcr\noalign{\kern-1pt\nointerlineskip} $\hfil\displaystyle{#1}\hfil$\crcr}}} \def\dt#1{{\buildrel {\hbox{\LARGE .}} \over {#1}}} \def\dtt#1{{\buildrel \bullet \over {#1}}} \def\der#1{{\pa \over \pa {#1}}} \def\fder#1{{\delta \over \delta {#1}}} \def\partder#1#2{{\partial #1\over\partial #2}} \def\parvar#1#2{{\delta #1\over \delta #2}} \def\secder#1#2#3{{\partial^2 #1\over\partial #2 \partial #3}} \def\on#1#2{\mathop{\null#2}\limits^{#1}} \def\bvec#1{\on\leftarrow{#1}} \def\oover#1{\on\circ{#1}} \def\begin{displaymath}{\begin{displaymath}} \def\end{displaymath}{\end{displaymath}} \def\nonumber{\nonumber} \def\pl#1#2#3{Phys.~Lett.~{\bf B {#1}} (19{#2}) #3} \def\np#1#2#3{Nucl.~Phys.~{\bf B {#1}} (19{#2}) #3} \def\prl#1#2#3{Phys.~Rev.~Lett.~{\bf #1} (19{#2}) #3} \def\pr#1#2#3{Phys.~Rev.~{\bf D {#1}} (19{#2}) #3} \def\cqg#1#2#3{Class.~and Quantum Grav.~{\bf {#1}} (19{#2}) #3} \def\cmp#1#2#3{Commun.~Math.~Phys.~{\bf {#1}} (19{#2}) #3} \def\jmp#1#2#3{J.~Math.~Phys.~{\bf {#1}} (19{#2}) #3} \def\ap#1#2#3{Ann.~of Phys.~{\bf {#1}} (19{#2}) #3} \def\prep#1#2#3{Phys.~Rep.~{\bf {#1}C} (19{#2}) #3} \def\ptp#1#2#3{Progr.~Theor.~Phys.~{\bf {#1}} (19{#2}) #3} \def\ijmp#1#2#3{Int.~J.~Mod.~Phys.~{\bf A {#1}} (19{#2}) #3} \def\mpl#1#2#3{Mod.~Phys.~Lett.~{\bf A {#1}} (19{#2}) #3} \def\nc#1#2#3{Nuovo Cim.~{\bf {#1}} (19{#2}) #3} \def\ibid#1#2#3{{\it ibid.}~{\bf {#1}} (19{#2}) #3} \newcommand{\newcommand}{\newcommand} \newcommand{\ra}{\rightarrow} \newcommand{\lra}{\leftrightarrow} \newcommand{\beq}{\begin{equation}} \newcommand{\eeq}{\end{equation}} \newcommand{\bea}{\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}} \def\epsilon{\epsilon} \def\lambda{\lambda} \def\frac{\frac} \def\nu_{e}{\nu_{e}} \def\nu_{\mu}{\nu_{\mu}} \def\nu_{\tau}{\nu_{\tau}} \def\sqrt{2}{\sqrt{2}} \def\rightarrow{\rightarrow} \newcommand{\sm}{Standard Model} \newcommand{\smd}{Standard Model} \newcommand{\barr}{\begin{eqnarray}} \newcommand{\earr}{\end{eqnarray}} \newcommand{\ccaption}[2]{ \begin{center} \parbox{0.85\textwidth}{ \caption[#1]{\small{\it{#2}}} } \end{center} } \begin{document} \begin{flushright} hep-ph/9904279 \\ CERN-TH/99-87 \\ \end{flushright} \vspace{15mm} \begin{center} {\LARGE \bf Can Neutrinos be Degenerate in Mass? } \\ \end{center} \vspace*{0.7cm} \begin{center} {\large {\bf John Ellis} \footnote{ {\tt email: <EMAIL>}} and {\bf Smaragda Lola} \footnote{ {\tt email: <EMAIL>}} } \end{center} \vspace*{0.1cm} \begin{center} {\large Theory Division, CERN, CH 1211 Geneva 23, Switzerland} \end{center} \vspace{1 cm} \begin{center} {\bf ABSTRACT} \end{center} {{\small We reconsider the possibility that the masses of the three light neutrinos of the Standard Model might be almost degenerate and close to the present upper limits from Tritium $\beta$ decay and cosmology. In such a scenario, the cancellations required by the latest upper limit on neutrinoless double-$\beta$ decay enforce near-maximal mixing that may be compatible only with the vacuum-oscillation scenario for solar neutrinos. We argue that the mixing angles yielded by degenerate neutrino mass-matrix textures are not in general stable under small perturbations. We evaluate within the MSSM the generation-dependent one-loop renormalization of neutrino mass-matrix textures that yielded degenerate masses and large mixing at the tree level. We find that $m_{\nu_e} > m_{\nu_\mu} > m_{\nu_\tau}$ after renormalization, excluding MSW effects on solar neutrinos. We verify that bimaximal mixing is not stable, and show that the renormalized masses and mixing angles are not compatible with all the experimental constraints, even for $\tan \beta$ as low as unity. These results hold whether the neutrino masses are generated by a see-saw mechanism with heavy neutrinos weighing $\sim 10^{13}$~GeV or by non-renormalizable interactions at a scale $\sim 10^5$~GeV. We also comment on the corresponding renormalization effects in the minimal Standard Model, in which $m_{\nu_e} < m_{\nu_\mu} < m_{\nu_\tau}$. Although a solar MSW effect is now possible, the perturbed neutrino masses and mixings are still not compatible with atmospheric- and solar-neutrino data.}} \noindent \thispagestyle{empty} \setcounter{page}{0} \vfill\eject \section{Introduction} Observations of solar and atmospheric neutrinos provide strong indications that neutrinos may oscillate between eigenstates with different masses $m_{\nu_i}$ \cite{SKam,KamMac}. The indications from solar neutrinos are for a difference in mass squared $\Delta m^2_{solar} \sim 10^{-4}$~\cite{MSW} to $10^{-10}$~eV$^2$ \cite{solar}, whereas the atmospheric neutrino data favour $\Delta m^2_{atmo} \sim 10^{-2}$ to $10^{-3}$~eV$^2$ \cite{SKam}. As is well known, oscillation experiments are not able to set the overall scale of the neutrino masses. However, there are some upper limits on these: astrophysical and cosmological constraints on dark matter suggest that $\Sigma_i m_{\nu_i} <$ few~eV \cite{Lyalpha}, and experiments on the endpoint of the Tritium $\beta$-decay spectrum suggest that $m_{\nu_i} < 2.5$~eV~\cite{Lobashev} for any mass eigenstate with a substantial electron flavour component~\cite{Barger}. The question then arises whether there are any indirect arguments bearing on the possibility that the three light neutrino masses might be approximately degenerate \cite{ol}-\cite{nonab} and close to these upper limits. Extending previous arguments in~\cite{GG} to include a new upper limit on neutrinoless double-$\beta$ decay~\cite{KKG}, we argue in the next section that, within the context of such degenerate neutrinos, this may already exclude the large-angle Mikheyev-Smirnov-Wolfenstein (MSW)~\cite{MSW} solution to the solar-neutrino problem. In this case, one would be forced into the vacuum-oscillation solution, and hence extreme mass degeneracy to one part in $10^{10}$. In the case of the large-angle MSW solution, the degeneracy would need to be to one part in $10^4$ or more. We subsequently discuss general features of the one-loop renormalization of neutrino mass matrices, using as an example one specific degenerate mass-matrix texture that accommodates neutrinoless double-$\beta$ decay via bimaximal mixing \cite{new1,GG}. We argue that this and other degenerate textures are generically unstable with respect to small perturbations, so that mixing does not remain bimaximal as suggested by oscillation data and the neutrinoless double-$\beta$ decay constraint. We evaluate within the Minimal Supersymmetric extension of the Standard Model (MSSM) the one-loop renormalization-group corrections to mass degeneracy in scenarios where the masses are generated either by a see-saw mechanism \cite{seesaw} at some high scale $M_N$ close to $M_{GUT}$, or by an effective operator at some low scale $\Lambda$ close to $M_{SUSY}$. We find that these corrections are significant, even for relatively small values of $\tan\beta$, and lead to the ordering $m_{\nu_e} > m_{\nu_\mu} > m_{\nu_\tau}$. In the case that neutrino masses of order ${\cal O} (2)$~eV arise via the see-saw mechanism, these corrections indicate that degenerate neutrinos are not compatible with all constraints, even for tan$\beta$ as low as 1. In particular, we note that the ordering of neutrino masses is {\it incompatible} with MSW solutions, because they require $m_{\nu_e} < m_{\nu_\mu}, m_{\nu_\tau}$. If neutrino masses arise from non-renormalizable interactions at a scale $\Lambda$ as low as 10-100 $m_{SUSY}$, the effects are smaller. Still, even in this case, we cannot obtain the required degeneracy for the vacuum oscillations, and MSW oscillations cannot be obtained. Finally, we discuss briefly renormalization effects in the minimal Standard Model. Their expected magnitude is similar to that in the MSSM in the low-$\tan\beta$ regime, since for a single Higgs field all the quark and charged-lepton mass hierarchies have to arise purely from the Yukawa couplings, and hence the $\tau$ coupling is small. However, in this case the sign of the Yukawa renormalization effects is opposite, tending to increase the magnitudes of the entries in $m_{eff}$ and leading to $m_{\nu_e} < m_{\nu_\mu}, m_{\nu_\tau}$, as required by the MSW mechanism. However, bimaximal mixing again cannot be maintained, so that this framework also appears incompatible with the combined constraints from neutrinoless double-$\beta$ decay and neutrino oscillation data. This analysis shows that schemes with degenerate neutrinos are very problematic, contrary to solutions with large neutrino hierarchies. Since the former may be obtained from non-Abelian symmetry structures \cite{nonab} and the latter from Abelian ones~\cite{abel,LLR,Grah}, renormalization-group effects on the neutrino mass eigenvalues may be providing important information about the underlying flavour structure of the fundamental theory. \section{Neutrinoless Double-$\beta$ Constraints on Degenerate Neutrinos \label{sec:neutr}} It was pointed out by Georgi and Glashow~\cite{GG} that, if the neutrino masses are close to the Tritium and cosmological upper limit so that relic neutrinos contribute at least one percent of the critical density of the Universe, then the upper limits on neutrinoless double-$\beta$ decay require the mixing angle for solar neutrino oscillations to be almost maximal. This is because the neutrinoless double-$\beta$ decay limit constrains the $ee$ component of the Majorana neutrino mass matrix in the charged-lepton flavour basis~\cite{GG}: \bea m_{eff}^{ee} \equiv \big\vert m_1\,c_2^2c_3^2e^{i\phi} + m_2\,c_2^2s_3^2e^{i\phi'} + m_3\, s_2^2\,e^{i2\delta} \big\vert < B \,. \label{doublebeta} \eea where $c_i,s_i$ denote $\cos\theta_{i},\sin\theta_{i}$ in the conventional $3 \times 3$ mixing parametrization \bea \pmatrix{\nu_e \cr \nu_\mu\cr \nu_\tau\cr}= \pmatrix{c_2c_3 & c_2s_3 & s_2e^{-i\delta}\cr -c_1s_3-s_1s_2c_3e^{i\delta} & +c_1c_3-s_1s_2s_3e^{i\delta} & s_1c_2\cr +s_1s_3-c_1s_2c_3e^{i\delta} & -s_1c_3-c_1s_2s_3e^{i\delta} & c_1c_2\cr}\, \pmatrix{\nu_1\cr \nu_2\cr \nu_3\cr}\,, \label{param} \eea where the diagonal matrix $m_{eff}^{diag}$ is $diag(m_1e^{i\phi}, m_2e^{i\phi'}, m_3)$, $\phi$ and $\phi'$ are phases in the light Majorana mass matrix, and $B$ is the experimental upper bound on $m_{eff}^{ee}$. In schemes with degenerate neutrinos, the differences between the mass eigenvalues $m_i$ in (\ref{doublebeta}) may be neglected~\cite{GG}. Moreover, given the upper limits on atmospheric oscillations into electron neutrinos established by Chooz \cite{chooz} and Super-Kamiokande \cite{SKam} , we follow~\cite{GG} and set $\theta_2 \approx 0$. Thus (\ref{doublebeta}) may be simplified to the form: \bea \big\vert \cos^2{\theta_3}\,e^{i\phi} + \sin^2{\theta_3}\,e^{i\phi'} \big\vert < {B \over {\overline m} } \label{simple} \eea where ${\overline m} \approx 2$~eV is the conjectured common mass scale of the (almost-)degenerate neutrinos. At the time of~\cite{GG}, the best available upper limit was $B < 0.46$~eV, and the constraint (\ref{simple}) could be satisfied for $\phi+\phi'\simeq \pi$ and $\vert\cos{2\theta_3}\vert < 0.23$, leading to $\sin^2{2\theta_3}> 0.95$. This was incompatible with the small-angle MSW solution for the solar neutrino data, but consistent with either the large-angle MSW solution or the vacuum-oscillation solution. Recently, however, a new upper limit $B < 0.2$~eV has been given~\cite{KKG}, from which we infer $\vert\cos{2\theta_3}\vert < 0.1$ and hence \beq \sin^2{2\theta_3}> 0.99. \label{KKG} \eeq Such maximal mixing is favoured by the vacuum-oscillation solution, but is disfavoured in the large-angle MSW solution, because $\sin^2{2\theta_3} = 1$ would yield an energy-independent suppression of all solar neutrinos~\cite{Giunti}. This disagrees with the Homestake data by at least three standard deviations, although the question persists whether the data could tolerate the lower limit in (\ref{KKG}). Several global fits to the solar-neutrino data have been published, including one by the Super-Kamiokande collaboration~\cite{SKdn} that takes into account its recent measurements of the day-night effect and yields $\sin^2{2\theta_3} < 0.99$~\footnote{If the day-night effect were not included, the upper limit would be $\sin^2{2\theta_3} < 0.95$: the magnitude of the day-night effect improves the quality of the fit in this large-angle MSW region~\cite{SKdn}.}. Other global fits often give smaller upper limits on $\sin^2{2\theta_3}$ \cite{SKdn}. We infer, provisionally, that the large-angle MSW solution is now excluded by neutrinoless double-$\beta$ decay if the neutrinos are near-degenerate, forcing us into the vacuum-oscillation solution, in which case the neutrino mass degeneracy must be at the level of one part in $10^{10}$. We could, however, imagine possible ways to evade this conclusion. Perhaps a small but non-trivial admixture of $\nu_{\mu} - \nu_e$ atmospheric oscillations could soften the bound (\ref{KKG}), and/or perhaps the solar-neutrino data could be stretched to accommodate it: $\sin^2{2\theta_3} = 1$ is excluded just at the $99.8 \%$ confidence level. Alternatively, the constraint (\ref{KKG}) would be weakened for degenerate neutrinos weighing less than 2~eV. We comment later how our conclusions would be affected if the large-angle MSW solution could be tolerated, in which case the neutrino mass degeneracy need only be to one part in $10^4$. Using the experimental information then available, a specific effective neutrino mass texture was proposed in~\cite{new1,GG}: \bea {m_{eff}} \propto \,\pmatrix{ 0 & {1\over\sqrt2} & {1\over\sqrt2}\cr {1\over\sqrt2} & {1\over2} & -{1\over2}\cr {1\over\sqrt2} & -{1\over2} & {1\over2}\cr } \label{GGtexture} \eea in the flavour basis where charged-lepton masses are diagonal, leading to bimaximal mixing. Other neutrino mass textures might be considered, depending on the theoretical assumptions and other phenomenological choices made. In the following, we shall use (\ref{GGtexture}) as an example, but frame our discussion in terms sufficiently general that it could be extended to other model textures. Any such texture can only be regarded as a first approximation, that might be modified by higher-order effects. These could include the possible contributions of higher-dimensional non-renormalizable operators. The above discussion suggests that any such contributions should change the mass eigenstates by at most one part in $10^{10}$ (or $10^4$), which is considerably more delicate than the expected hierarchy $m_{GUT}/m_{P} \approx 10^{-2}$. In the absence of any detailed theory of such contributions, one cannot say that this is necessarily a problem. However, global symmetries are not normally expected to be exact at the Planck scale, so the mass-degeneracy constraint is potentially powerful. Moreover, the mixing angles in such degenerate mass-matrix models are inherently unstable when higher-order perturbations are switched on, as we discuss in more detail later. \section{Renormalization-Group Effects on Neutrino Mass Textures} Calculable and potentially significant breakings of the neutrino mass degeneracies are provided by renormalization-group effects. However, these depend on the specific neutrino model framework adopted. We consider here two possibilities: one is the conventional see-saw, with a singlet-neutrino mass scale $M_N \sim 10^{13}$~GeV \footnote{ We note that the heavy Majorana neutrino masses $M_N$ need not be degenerate. On the contrary, flavour symmetries indicate that should have a structure determined by the flavour charges of the $N$ fields, and of other singlet fields in a given model. However, we do not discuss explicitly here this structure, which could also in principle affect the amount of renormalization-group running.}, and the other is a model where the light Majorana neutrino masses are simply generated by a new non-renormalizable interaction, such as $\nu_L \nu_L H H$, at a mass scale $\Lambda \sim 10^5$~GeV, close to $m_{SUSY} = 10^3$~GeV. Below we give numerical results for both scenarios. In the see-saw case, between the GUT scale and the scale $M_{N}$ of the heavy Majorana neutrinos, there is an effect on the mixing angle due to the renormalization-group running of the Dirac neutrino coupling $Y_N$ \cite{VB}: \begin{eqnarray} 8\pi^2 {d\over dt}({ Y}_N { Y}_N^\dagger)& = & \{ -\sum_i c_N^i g_i^2+3 ({ Y}_N { Y}_N^\dagger) +{\rm Tr}[3({ Y}_U { Y}_U^\dagger)+({ Y}_N { Y}_N^\dagger)] \} ({ Y}_N { Y}_N^\dagger) \nonumber \\ & &\qquad \qquad +{1\over 2}\{({ Y}_E { Y}_E^\dagger) ({ Y}_N { Y}_N^\dagger)+({ Y}_N { Y}_N^\dagger) ({ Y}_E { Y}_E^\dagger)\} \label{firstRG} \end{eqnarray} Here and subsequently we work at the one-loop level, and denote the renormalization-group scale by $t \equiv \ln{\mu}$. In the MSSM, $c^i_N = (3/5, 3, 0)$, and we denote the Dirac couplings of other types of fermion $F$ by $Y_F$. It is apparent from (\ref{firstRG}) that large Yukawa couplings have a bigger effect on $m^{D}_{33}$ than on the rest of the mass-matrix elements, and tend in general to lower $Y_N$. This alters the structure of the Dirac mass matrix, in turn affecting the magnitudes of the mixing angles. These effects become more relevant in examples where cancellations between various entries may lead to amplified mixing in $m_{eff}$. However, we assume here that any neutrino-mass texture~\cite{GG} is defined at the characteristic scale $M_N$ of the see-saw mechanism. Therefore, in the current discussion we use this first part of the run only in order to define the initial conditions for the gauge and Yukawa couplings at $M_N$, but not to modify the neutrino-mass texture. Since the exact form of $m_{eff}$ depends on the right-handed Majorana mass matrix, we simply assume that this has the form that is required in order to lead to a specific texture at $M_N$. Having set the initial conditions for $g_i$ and $\lambda_i$ at $M_N$, we note that ${ Y}_N$ decouples below the right-handed Majorana-mass scale, where the relevant running is that of the effective neutrino-mass operator \cite{Bab,run}: \begin{equation} 8\pi^2 {d\over dt}{ m_{eff}} = \{-({3\over 5} g_1^2+3 g_2^2)+ {\rm Tr} [3{ Y}_U { Y}_U^\dagger]\}m_{eff} +{1\over 2}\{({ Y}_E { Y}_E^\dagger)m_{eff} + m_{eff} ({ Y}_E { Y}_E^\dagger)^T\} \ , \label{MASSES} \end{equation} This is the basic equation for the running of the various entries of the effective light-neutrino mass matrix. We continue to work in the basis where the charged-lepton mass matrix is diagonal, which is also the basis in which the neutrino-mass texture is specified. In previous work \cite{ELLN2}, we discussed the running of the (23) mixing angle, but here we focus more on the evolution of the $m_{eff}$ entries themselves, extending our previous discussion to encompass solutions to the solar neutrino problem. For this purpose, we use the following differential equations for individual elements of the effective neutrino-mass matrix: \bea \frac{1}{m_{eff}^{ij}}\frac{d}{d t}m_{eff}^{ij}&=& \frac{1}{8\pi^2}\left( -c_i g_i^2 + 3 h_t^2 + \frac{1}{2} ( h_{i}^2 + h_j^2) \right) \label{RGmeff} \eea It is convenient for the subsequent discussion to define the integrals \bea I_g& =& exp[\frac{1}{8\pi^2}\int_{t_0}^t(-c_i g_i^2 dt)]\\ I_t &=& exp[\frac{1}{8\pi^2}\int_{t_0}^t h_t^2 dt]\\ I_{h_i} &=& exp[\frac 1{8\pi^2}\int_{t_0}^t h_{i}^2 dt] \eea where in $I_{h_i}$ and $h_i$ the subindex $i$ refers to the charged-lepton flavours $e, \mu$ and $\tau$. Simple integration of (\ref{RGmeff}) yields \bea \frac{m_{eff}^{ij}}{m_{eff,0}^{ij}}&=& exp\left\{ \frac{1}{8\pi^2}\int_{t_0}^t \left(-c_i g_i^2 + 3 h_t^2 + \frac{1}{2} ( h_{i}^2 + h_j^2) \right)\right\} \nonumber\\ &=& I_g\cdot I_t \cdot \sqrt{I_{h_i}} \cdot \sqrt{I_{h_j}} \label{RGis} \eea where the initial conditions are denoted by $ m_{eff,0}^{ij}$. As we have already noted, these conditions are defined at $M_N$, the scale where the neutrino Dirac coupling $h_N$ decouples from the renormalisation-group equations. Using (\ref{RGis}), we see that an initial texture $m_{eff,0}^{ij}$ at $M_N$ is modified to become \bea m_{eff} = \left ( \begin{array}{ccc} m_{eff,0}^{11} ~I_e & m_{eff,0}^{12} ~\sqrt{I_\mu} ~\sqrt{I_e} & m_{eff,0}^{13} ~\sqrt{I_e} ~\sqrt{I_\tau} \\ & & \\ m_{eff,0}^{21} ~\sqrt{I_\mu} ~\sqrt{I_e} & m_{eff,0}^{22} ~I_\mu & m_{eff,0}^{23} ~\sqrt{I_\mu} ~\sqrt{I_\tau} \\ & & \\ m_{eff,0}^{31} ~\sqrt{I_e} ~\sqrt{I_\tau} & m_{eff,0}^{32} ~\sqrt{I_\mu} ~\sqrt{I_\tau} & m_{eff,0}^{33} ~I_\tau \end{array} \right) \nonumber \\ = \left ( \begin{array}{ccc} \sqrt{I_e} & 0 & 0 \\ 0 & \sqrt{I_\mu} & 0 \\ 0 & 0 & \sqrt{I_\tau} \end{array} \right) \cdot \left ( \begin{array}{ccc} m_{eff,0}^{11} & m_{eff,0}^{12} & m_{eff,0}^{13} \\ & & \\ m_{eff,0}^{21} & m_{eff,0}^{22} & m_{eff,0}^{23} \\ & & \\ m_{eff,0}^{31} & m_{eff,0}^{32} & m_{eff,0}^{33} \end{array} \right) \cdot \left ( \begin{array}{ccc} \sqrt{I_e} & 0 & 0 \\ 0 & \sqrt{I_\mu} & 0 \\ 0 & 0 & \sqrt{I_\tau} \end{array} \right) \label{factor} \eea at $m_{SUSY}$. We evaluate subsequently the integrals $I_{e,\mu,\tau}$ appearing in this renormalization. However, we can already extract some important qualitative information from (\ref{factor}). $\bullet$ We first note that because of the factorization in (\ref{factor}), although the individual masses and mixings get modified, any mass matrix which is singular with a vanishing determinant - leading to a zero mass eigenvalue - remains so at the one-loop level. However, one should expect modifications at the two-loop level, which might be an interesting mechanism for generating a non-trivial but large neutrino-mass hierarchy. $\bullet$ The Yukawa renormalization factors $I_i$ are less than unity, and lead to the mass ordering $m_{\nu_e} > m_{\nu_\mu} > m_{\nu_\tau}$, to the extent that such naive flavour identifications are possible. $\bullet$ One would expect that for values of $I_{\tau}$ substantially different from unity - which occur for large $\tan\beta$ in particular~\footnote{Most flavour-symmetry models in the literature assume large $\tan\beta$.} - the renormalization effects on the (23) sector would be especially significant. However, there can be important effects even in the first-generation sector. These can be significant for two reasons. One is that, in view of the neutrinoless double-$\beta$ decay analysis given above, very small mass differences may be required for addressing the solar neutrino problem, so we should keep even small renormalization effects in mind. The other is that, when off-diagonal entries in $m_{eff,0}^{ij}$ are large as in the sample texture (\ref{GGtexture}), the $I_{\tau}$ renormalization effects feed through into all differences in mass eigenvalues. $\bullet$ In the case of the mixing angles, we recall that renormalization effects may either enhance or suppress the mixing. In particular, it has been noted in the case that $m_{eff,0}^{22} = m_{eff,0}^{33}$ and atmospheric-neutrino mixing is maximal somewhere above the electroweak scale, the maximal mixing may not survive down to low energies if the $\tau$ Yukawa coupling is large, at least for certain textures, depending on the magnitude of the (23) entries. To illustrate this, we will make some generic comments on the case of $2 \times 2$ mixing, and we will return to neutrino-mixing effects for the texture (\ref{GGtexture}) in the next section. In the case of simple $ 2\times 2$ mixing, we see from \begin{equation} \sin^2 2\theta_{23} = \frac{4 (m_{eff,0}^{23}) ^2}{( m_{eff,0}^{33} - m_{eff,0}^{22})^2 + 4 (m_{eff,0}^{23})^2} \end{equation} that the degeneracy between $m_{eff}^{22}$ and $m_{eff}^{33}$ becomes important only if $m_{eff}^{23}$ is of the same order as $m_{eff}^{33} - m_{eff}^{22}$. It is known that large neutrino-mass hierarchies can be generated by two-generation textures of the following forms \cite{recent,ELLN2} in the basis where the charged leptons are diagonal: \bea \left ( \begin{array}{cc} x^2 & x \\ x & 1 \end{array} \right ), ~ \left ( \begin{array}{cc} 1 & \pm 1 \\ \pm 1 & 1 \end{array} \right ), ~ \left ( \begin{array}{cc} 1 & 1 \\ -1 & -1 \end{array} \right ), ~ \left ( \begin{array}{cc} 1 & x \\ x & 1 \end{array} \right ) \label{tex1} \eea where the first solution has a large but non-maximal mixing in contrast with the others. For the first texture \cite{Grah}, where $m_{eff,0}^{22} < m_{eff,0}^{33}$, the renormalization-group effects on the mixing clearly are negligible, and the same is also true for the third texture, due to the signs of the entries~\footnote{The renormalization of three-generation textures is more complicated, as we see in the next section.}. In the case of the second texture, one expects mild changes despite the fact that $m_{eff,0}^{22} = m_{eff,0}^{33}$, because $m_{eff,0}^{23} $ is large. In the fourth texture \cite{degen1}, $m_{eff,0}^{22} = m_{eff,0}^{33}$, and $m_{eff,0}^{23}$ is small, which is exactly the type of solution that is very unstable under renormalization-group running. Finally, we note that solutions with small hierarchies and large mixing \cite{degen2} of the type \bea \left ( \begin{array}{cc} x & 1 \\ 1 & x' \end{array} \right ) \label{tex2} \eea are also expected to be stable under the renormalization group. To complete this algebraic discussion of renormalization-group effects, we now comment on the second case of interest, in which the neutrino-mass texture is assumed to be generated by non-renormalizable interactions at some relatively low mass scale $\Lambda \sim 10^5$~GeV, such as $\nu_L \nu_L H H / \Lambda$. In this case, there are no neutrino Dirac couplings to be renormalized, so the renormalization-group running between $M_{GUT}$ and $\Lambda$ is the same as equations (\ref{MASSES}) to (\ref{factor}), with Yukawa couplings only for quarks and charged leptons. Below the scale $\Lambda$, $m_{eff}$ runs in the same way as we discussed previously below $M_N$, but the range of scales over which the renormalization must be computed is greatly reduced. \section{Numerical Results} We now present some numbers for $I_\tau$ and $I_\mu$, in order to exemplify renormalization effects on the textures. We take as illustrative initial conditions~\footnote{Although the runnings of the gauge couplings and of $h_t$ factor out, they nevertheless affect the magnitude of $h_\tau$ and hence the exact value of $I_\tau$ that one derives.} $\alpha_{GUT}^{-1} = 25.64$, $M_{GUT} = 1.1 \cdot 10^{16}$~GeV and $m_{SUSY} = 1$ TeV. We also take $h_t = 3.0$, and choose $h_b / h_\tau$ such that an intermediate scale $M_N$ is consistent with the observed pattern of fermion masses \footnote{The choice of input parameters needed to reproduce exactly the observed fermion masses depends on $\tan\beta$, but incorporating this refinement is unnecessary for our purposes. We comment only that, for small $\tan\beta$, an intermediate scale $M_N$ may be consistent with the values of $m_b$ and $m_\tau$ measured at low energies, at the cost of a certain deviation from bottom--tau mass unification \cite{VB}, which may be $\sim 10 \%$ for $M_N \approx 10^{13}~{\rm GeV}$. However, this may be corrected~\cite{LLR}, if there is sufficient mixing in the charged-lepton sector. For completeness, we note that we use $h_N = 3.0$: this choice has a small impact on the initial conditions at the scale $M_N$.}. We use the physical $\mu$ and $e$ masses to fix $h_\mu, h_e$. Finally, we take $M_N = 10^{13}$~GeV as our default, mentioning later the effects with a different choice. The values of $I_\tau$ and $I_\mu$ that we find with these inputs are given in the first three columns of Table~1 and plotted in Fig.~1. These results may be used to estimate the effects on the neutrino eigenvalues, mixings and mass differences in the specific texture (\ref{GGtexture}), as shown in the last three columns of Table~1 and in Fig.~2. \begin{table}[tbp] \begin{center} \begin{small} \begin{tabular}{|c|c|c|c|c|c|} \hline \hline $h_\tau$ & $I_\tau$ & $I_\mu$ & $m_{3}$ & $m_{2}$ & $m_{1}$ \\ \hline 3.0 & 0.826 & 0.9955 & 0.866 & -0.952 & 0.997 \\ \hline 1.2 & 0.873 & 0.9981 & 0.903 & -0.966 & 0.998 \\ \hline 0.48 & 0.9497 & 0.9994 & 0.962 & -0.987 & 0.9996 \\ \hline 0.10 & 0.997 & 0.99997 & 0.9478 & -0.9993 & 0.99998 \\ \hline 0.013 & 0.99997 & 1.00000 & 0.99998 & -0.99999 & 1.00000 \\ \hline \hline \end{tabular} \end{small} \end{center} \caption{ {\it Values of $I_\tau$ and $I_\mu$, for $M_{N} = 10^{13}$~GeV and different choices of $h_\tau$. Also tabulated are the three renormalized mass eigenvalues calculated from the sample texture (\ref{GGtexture}). }} \end{table} We see that the renormalization-group effects on the neutrino-mass eigenvalues are significant. Since they are larger for the second- and third-generation leptons, as already commented, we find the following ordering of the light neutrino masses for textures with degenerate eigenvalues at $M_N$: $m_{\nu_{e}} > m_{\nu_{\mu}} > m_{\nu_{\tau}}$, which is {\it incompatible} with MSW solutions to the solar-neutrino problem. It is apparent from Table~1 and Fig.~2 that the breaking of the neutrino-mass degeneracy in this model is unacceptable for any value of $h_\tau$ corresponding to $1 < {\rm tan} \beta < 58$. \begin{figure}[tbp] \vspace*{-4.4 cm} \centerline{\epsfig{figure=Fig1.ps,width=1.2\textwidth,clip=}} \vspace*{-13.0 cm} \caption{{\it Numerical values of $I_{\mu}$ and $I_\tau$ for different initial values of $h_\tau$, assuming $M_N = 10^{13}$~GeV. The values of $h_\tau^0 = $ 0.013, 0.03, 0.05 and 0.5 to 3 in steps of 0.5 correspond to tan$\beta$ 1, 3.8, 6.5, 43.8, 53.6, 56.3, 57.4, 57.9 and 58.2, respectively. }} \vspace*{-2.4 cm} \centerline{\epsfig{figure=Fig2.ps,width=1.2\textwidth,clip=}} \vspace*{-13.0 cm} \caption{ {\it Renormalization of $m_{eff}$ eigenvalues for different initial values of $h_\tau$ corresponding to values of tan$\beta$ in the range 1 to 58, assuming the particular neutrino-mass texture (\ref{GGtexture}) and $M_N = 10^{13}$~GeV. We see that the vacuum-oscillation scenario is never accommodated.}} \end{figure} We now discuss the renormalization of the neutrino mixing angles, using as a particular example the texture (\ref{GGtexture}). Initially we make a generic discussion, based on analytic formulas, and then we illustrate the discussion using two numeric examples. We consider the following parametrization of a perturbation from the initial texture, motivated by the structure (\ref{factor}): \bea {m_{eff}'} \propto \, \left ( \begin{array}{ccc} 0 & {1\over\sqrt2} & {1\over\sqrt2} (1+{\epsilon \over 2}) \\ & & \\ {1\over\sqrt2} & {1\over2} & -{1\over2} (1+{\epsilon \over 2}) \\ & & \\ {1\over\sqrt2} (1+{\epsilon \over 2}) & -{1\over2} (1+{\epsilon \over 2}) & {1\over2} (1+{\epsilon }) \end{array} \right) \label{GGtextureP} \eea where $\epsilon$ is a small quantity, which might arise from renormalisation group running or from some other higher-order effects such as higher-dimensional non-renormalizable operators. This perturbation lifts the degeneracy of the eigenvalues, which are now given by \bea 1, ~~ -1 - {\epsilon \over 4}, ~~1 + {3 \epsilon \over 4} \nonumber \eea To this order, the eigenvectors are independent of $\epsilon$ and given by \bea V_1 = \left ( \begin{array}{c} {1 \over \sqrt{3}} \\ \sqrt{ \frac{2}{3} } \\ 0 \end{array} \right ),~~ V_2 = \left ( \begin{array}{c} {1 \over \sqrt{2}} \\ -{1 \over 2} \\ -{1 \over 2} \end{array} \right ), ~~ V_3 = \left ( \begin{array}{c} {1 \over \sqrt{6}} \\ -{1 \over {2 \sqrt{3}}} \\ {\sqrt{3}\over {2}} \\ \end{array} \right ) \label{vec1} \eea so that the mixing expected in this type of texture does not depend on $\epsilon$, as long as it is non-zero. However, this mixing is {\it not} bimaximal. The vectors (\ref{vec1}) are also eigenvactors of the unrenormalised texture (\ref{GGtexture}). Since this unperturbed texture has two exactly degenerate eigenvalues, there is arbitrariness in the choice of eigenvectors: the vectors corresponding to the two degenerate eigenvalues can be rotated to different linear combinations, which are still eigenvectors of the neutrino mass matrix and still obey the orthogonality conditions. One example is the choice \bea V_1 & = & \frac{1}{\sqrt{3}} V_1' + \sqrt{\frac{2}{3}} V_3' \nonumber \\ V_3 & = & \frac{1}{\sqrt{3}} V_3' -\sqrt{\frac{2}{3}} V_1' \nonumber \eea which gives \bea V_1' = \left ( \begin{array}{c} 0 \\ \frac{1}{\sqrt{2}} \\ -\frac{1}{\sqrt{2}} \end{array} \right ),~~ V_2' = \left ( \begin{array}{c} {1 \over \sqrt{2}} \\ -{1 \over 2} \\ -{1 \over 2} \end{array} \right ), ~~ V_3' = \left ( \begin{array}{c} \frac{1}{\sqrt{2}} \\ \frac{1}{2} \\ \frac{1}{2} \end{array} \right ) \label{vec2} \eea corresponding to bimaximal mixing: $\phi_1 = {\pi \over 4}$, $\phi_2 = 0$ and $\phi_3 = {\pi \over 4}$. However, one cannot in general expect this combination of eigenvectors to be stable when the degenerate texture is perturbed, and the above analysis shows that, indeed, it is not. On the contrary, it is the direction given by (\ref{vec1}) that is stable, and the absence of the parameter $\epsilon$ in the eigenvectors indicates that we may expect only minor modifications in the mixing, for $\tau$ couplings in the range $3.0-0.013$. We illustrate the instability of bimaximal mixing and the stability of the eigenvectors (\ref{vec1}) with a numerical analysis of two extreme cases with $h_{\tau}^0 = 3$ and $ 0.013$. Using the values of $I_{\tau ,\mu}$ given in Table 1, we determine the full renormalized mass matrices to be: \bea m_{eff}^{1,ren} = \left ( \begin{array}{ccc} 0 & 0.705 & 0.64 \\ 0.705 & 0.497 & -0.45 \\ 0.64 & -0.45 & 0.41 \end{array} \right), ~~ m_{eff}^{2,ren} = \left ( \begin{array}{ccc} 0 & 0.7071 & 0.7071 \\ 0.7071 & 0.5 & -0.499992 \\ 0.7071 & -0.499992 & 0.499985 \end{array} \right), ~~ \eea respectively, which are to be compared with the initial form (\ref{GGtexture}) of the texture. Then, for \bea U_1 = \left ( \begin{array}{ccc} 0.5804 & 0.8143 & 0.0065 \\ -0.7075 & 0.5003 & 0.4992 \\ 0.4032 & -0.2943 & 0.8665 \end{array} \right), ~ U_2 = \left ( \begin{array}{ccc} 0.57864 & 0.815578 & 0.00274 \\ -0.7071 & 0.5 & 0.5 \\ 0.406418 & -0.29126 & 0.866021 \end{array} \right) \eea we find that \bea m_{eff,diag }^{1,ren} = U_1 \cdot m_{eff}^{1,ren} \cdot U_1^T = \left ( \begin{array}{ccc} 0.997 & 0 & 0 \\ 0 & -0.952 & 0 \\ 0 & 0 & 0.866 \end{array} \right) \eea and \bea m_{eff,diag }^{2,ren} = U_2 \cdot m_{eff}^{2,ren} \cdot U_2^T = \left ( \begin{array}{ccc} 1.00000 & 0 & 0 \\ 0 & -0.99999 & 0 \\ 0 & 0 & 0.99998 \end{array} \right) \eea reflecting the reverse mass ordering mentioned above. On the other hand, as we have already remarked, the eigenvectors (and thus the mixing matrices) are stable. It is easy to check that the matrices $U_{1,2}$ are unitary ones, and in the notation of (\ref{param}), correspond to the values \bea \phi_1 \approx -0.327, ~~\phi_2 \approx 0.415, ~~ \phi_3 \approx -0.884 \eea In our notation, atmospheric-neutrino mixing is controlled by the parameter $\phi_1$, for which we find $\sin^2 2\phi_1 \approx 0.37$, whereas solar-neutrino mixing is controlled by $\phi_3$, for which we find $\sin^2 2\phi_3 \approx 0.96$. We see therefore that even small perturbations of exact neutrino degeneracy cause large effects on the neutrino mixing angles, which then {\em conflict with the combined bounds from neutrinoless double-$\beta$ decay and oscillation data}. This example shows that the mixing differs significantly from that postulated in the unperturbed degenerate texture, an effect not visible in a naive $2 \times 2$ analysis. Up to now we have been discussing the situation where light neutrino masses arise through the see-saw mechanism, and therefore $m_{eff}$ arises at a scale $10^{13}$~GeV. However, if $m_{eff}$ arises at a significantly lower scale $\Lambda$, for example via effective operators of the form $\nu_L \nu_L H H / \Lambda$, as discussed at the end of the previous section, the integrals $I_{\tau}$ and $I_{\mu}$ are now much closer to unity. This happens because (i) the range where $m_{eff}$ runs is significantly decreased, and (ii) the starting value of $h_{\tau}$ at $\Lambda$ is also smaller, due to the run from $M_{GUT}$ to $\Lambda$ being over a relatively wide range. Shown in Table~2 and Fig.~3 are our calculations of $I_\tau$ and $I_\mu$, using the same parameters as in Table~1 and Fig.~1, except that now we run down from $\Lambda = 10^{5}$~GeV, corresponding to a scenario where $h_N$ is small or zero~\footnote{Since the logarithmic range of renormalization-group running is short in this case, finite renormalization effects may be relatively more significant than in the $M_N = 10^{13}$~GeV case. However, their evaluation requires detailed modelling of thresholds, which lies beyond the scope of this paper. We consider it unlikely that the qualitative conclusions of this paper would be affected by their inclusion.}. The effects on the eigenvalues appear in Table~2 and Fig.~4, from which we see that, although the mass ratios are now closer to unity than in the previous case, the effects of the running can still not be neglected, when compared to the small mass differences required by the solar neutrino data. We again find that the full range $1 < {\rm tan} \beta < 58$ is excluded. On the other hand, the effect on the mixing angle is similar to the previous case, since it is practically unchanged under small perturbations, as we discussed earlier. \begin{table}[h] \begin{center} \begin{small} \begin{tabular}{|c|c|c|c|c|c|} \hline \hline $h_\tau$ & $I_\tau$ & $I_\mu$ & $m_{3}$ & $m_{2}$ & $m_{1}$ \\ \hline 3.0 & 0.973 & 0.9988 & 0.9795 & -0.9929 & 0.9992 \\ \hline 1.2 & 0.975 & 0.9995 & 0.9815 & -0.9937 & 0.9996 \\ \hline 0.48 & 0.986 & 0.9998 & 0.9897 & -0.9965 & 0.9999 \\ \hline 0.10 & 0.9999 & 0.99999 & 0.9993 & -0.9997 & 0.99999 \\ \hline 0.013 & 0.99999 & 1.000000 & 0.999992 & -0.999997 & 1.00000 \\ \hline \hline \end{tabular} \end{small} \end{center} \caption{ {\it Values of $I_\tau$ and $I_\mu$, for $\Lambda = 10^{5}$~GeV and different choices of $h_\tau$. Also tabulated are the three renormalized mass eigenvalues calculated from the sample texture (\ref{GGtexture}). }} \end{table} \begin{figure}[tbp] \vspace*{-4.4 cm} \centerline{\epsfig{figure=Fig3.ps,width=1.2\textwidth,clip=}} \vspace*{-13.0 cm} \caption{{\it Numerical values of $I_{\mu}$ and $I_\tau$ for different initial values of $h_\tau$, assuming $\Lambda = 10^{5}$~GeV. The corresponding values of tan$\beta$ are roughly the same as in the previous case.}} \vspace*{-2.4 cm} \centerline{\epsfig{figure=Fig4.ps,width=1.2\textwidth,clip=}} \vspace*{-13.0 cm} \caption{ {\it Renormalization of $m_{eff}$ eigenvalues for different initial values of $h_\tau$ corresponding to values of tan$\beta$ in the range 1 to 58, assuming the particular neutrino-mass texture (\ref{GGtexture}) and $\Lambda = 10^{5}$~GeV. We see again that the vacuum-oscillation scenario is never accommodated.}} \end{figure} We comment finally on the renormalization-group effects in the minimal non-supersymmetric Standard Model. The evolution equation for $m_{eff}$ is \begin{equation} 16 \pi^2 {d m_{eff} \over d t} = ( - 3 g_2^2 + 2 \lambda + 2 S ) m_{eff} - {1\over 2} ( (m_{eff} (Y_e^\dagger Y_e) + (Y_e^\dagger Y_e)^T m_{eff}), \end{equation} where $\lambda$ is the Higgs coupling: $M_H^2 = \lambda v^2$, and $ S \equiv Tr( 3 Y_u^\dagger Y_u + 3 Y_d^\dagger Y_d + Y_e^\dagger Y_e ) $~\cite{Bab}. Although the running of $m_{eff}$ differs from the MSSM, the structure is similar. In particular, the contributions proportional to $g_2$, $\lambda$ and $S$ are the same for all entries, and thus the exponential factors that are obtained by integrating the renormalization-group equations multiply all entries, just as $I_t$ and $I_g$ did in the case of the MSSM. The term that affects the relative runnings of the various entries is again $ {1\over 2} ( (m_{eff} (Y_e^\dagger Y_e) + (Y_e^\dagger Y_e)^T m_{eff}$, though with a sign opposite from the MSSM, meaning that now the Yukawa couplings increase the entries in $m_{eff}$. An important feature in the Standard Model, since it has only one Higgs field, is that the mass hierarchies between fermions with opposite electroweak hypercharge have to arise purely from the Yukawa couplings. Hence the starting value of the $\tau$ coupling is small in this case, and therefore the effects are expected quantitatively to be similar to those in the low-$\tan\beta$ MSSM, but in the opposite direction. This is interesting, since whereas starting from degenerate-mass neutrinos in the MSSM we expect low-energy neutrino hierarchies of the type $m_{\nu_\tau} < m_{\nu_\mu} < m_{\nu_e} $, in the Standard Model we expect the opposite ordering of masses: $m_{\nu_\tau} > m_{\nu_\mu} > m_{\nu_e} $, which is the right sign for MSW solutions of the solar-neutrino problem. However, the SM case shares with the MSSM case the instability in the bimaximal mixing. This means that the renormalized mass matrix is again incompatible with the combined constraints from neutrinoless double-$\beta$ decay and oscillation data, even though the breaking of the mass degeneracy might appear compatible with the MSW solution to the solar-neutrino problem. \section{Conclusions} We have studied in this paper the circumstances under which neutrino masses can be degenerate and close to the present upper bounds from Tritium $\beta$ decay and astrophysics. We find that such schemes are severely constrained. In particular, the new upper limit on neutrinoless double-$\beta$ decay~\cite{KKG}, in combination with the rest of the solar-neutrino data, seems to exclude even the large-angle MSW solution to the solar-neutrino problem, and thus degenerate neutrinos may be compatible only with vacuum oscillations. However, in this case extreme mass degeneracy to one part in $10^{10}$ is required. Even if such a degeneracy is guaranteed by a symmetry at a certain scale, we find that renormalization group effects lift this degeneracy. In the MSSM with light neutrino masses arising through the see-saw mechanism, the effects on the eigenvalues are larger for larger $\tan\beta$, and have the wrong sign for MSW solutions. For a given $\tan\beta$, the effects are reduced if $m_{eff}$ arises via a non-renormalizable operator such as $\nu_L \nu_L H H / \Lambda$ at a significantly lower scale than the $10^{13}$~GeV required by the see-saw. Even in this case, however, the effects may not be neglected, in view of the extreme mass degeneracy that is required. Moreover, we find that even small perturbations shift the neutrino mixing angles by finite amounts, violating the combined constraints from neutrinoless double-$\beta$ decay and oscillation data. Finally, we find in the minimal Standard Model renormalization effects that are qualitatively similar to those of the low-$\tan\beta$ MSSM, but with opposite signs, thus leading to reversed low-energy neutrino-mass ordering. In this case, the large-angle MSW solution may survive, but the instability in the degenerate neutrino mixing angles means that the combined constraints from neutrinoless double-$\beta$ decay and oscillation data are still violated. Our analysis indicates that degenerate neutrino-mass textures have many problems when renormalization effects are taken into account. These results may provide hints on the appropriate framework for flavour symmetries, with Abelian models~\cite{abel,LLR,Grah} apparently favoured. \vspace*{0.35 cm} \noindent {\bf Acknowledgements:} We thank A. Casas, J.R. Espinosa, Y. Nir and S. Pakvasa, for comments on the paper.
\section{Introduction} The global star formation history of the universe holds clues to understanding the formation and evolution of galaxies. The recent detections of dust enshrouded galaxies at $z > 1$ at sub-millimeter wavelengths (Smail, Ivison \&\ Blain 1997; Hughes et al. 1998; Barger et al. 1998; Lilly et al. 1998) suggest that significant amounts of star formation activity at high redshifts may be obscured. UV-selected galaxies most likely represent only one segment of the population of active star-forming systems at early epochs. The difficulty in estimating the degree of extinction in the rest-frame UV introduces significant uncertainties in the star formation history (Pettini et al. 1998; Heckman et al. 1998; Blain et al. 1999; Malkan 1998). Little is known about the properties of normal galaxies in the region between $ 1 < z < 2$, where neither the 4000\AA\ break nor the Ly continuum break are easily accessible. The determinations of the global star formation history based on either the 2500\AA\ and 1500\AA\ continuum luminosities, or on the abundance of damped Ly$\alpha$ systems (e.g. Pei \& Fall 1995; Madau et al 1996; Steidel et al. 1998, Pei, Fall, \& Hauser 1999), imply either a peak or a plateau in the $1 < z < 2$ range. The near-IR offers one means of accessing both redshift indicators and measures of star formation at these redshifts. NICMOS offered a unique opportunity to perform slitless spectroscopy in the near-IR. The extremely low background achieved on HST at $\lambda < 1.9\mu$ allows for sensitive surveys for H$\alpha$ to $z = 1.9$. We carried out a survey of random fields with the slitless G141 grism ($\lambda_c = 1.5\mu$, $\Delta\lambda=0.8\mu$), covering a total $\sim 64$ square arc-minutes (McCarthy et al. 1999; hereafter paper I). Our survey has equal or greater depth than current ground-based narrow-band imaging programs, and due to its large wavelength coverage probes an order of magnitude more co-moving volume. The details of the survey and the spectra of the emission-line objects are given in paper I. In this $Letter$ we present the H$\alpha$ luminosity function at $ 0.7 < z < 1.9$ derived from the H$\alpha$ emission-line galaxy sample described in Paper I. We also discuss the implications of our results in the context of galaxy formation and evolution. Throughout this paper we use H$_0=50$~km s$^{-1}$ Mpc$^{-1}$ and q$_0=0.5$. \section{The H$\alpha$ Luminosity Function at $z \sim 0.7 - 1.9$} \medskip \subsection{The properties of the sample} The bandpass of the G141 grism in camera 3 of NICMOS limits the range of redshifts within which we can detect H$\alpha$ to $0.7 - 1.9$. The identification of the emission lines as H$\alpha$ is based on H-band apparent continuum magnitudes, their equivalent widths and the lack of other detected lines within the G141 bandpass. In two of the three cases for which optical spectroscopy has been attempted with LRIS on the Keck 10m telescope by Teplitz et al. (1999), the line identifications have been confirmed by detection of their [OII]3727 emission. These objects (J0055+8518a, $z=0.76$, J0622-0018a, $z=1.12$) are at the low redshift end of our sample and therefore are the most accessible to optical spectroscopy. In the third case (J1237+6219A, $z=1.37$) no emission or absorption features were detected in the spectrum. \bigskip \subsection{Calculation of the H$\alpha$ luminosity function at $z \sim 0.7 - 1.9$} We used the $\rm 1/V_{max}$ method (Schmidt 1968) to compute the luminosity function. Each galaxy is assigned a volume, $\rm V_{max}$, equal to the volume within which it could lie and be detected by our survey. We calculated the maximum co-moving volume for each galaxy in each field with its appropriate line flux limit. The maximum volume is defined as \begin{equation} V_{max} = \Omega \times \int_{zmin}^{zmax} \left({c \over H_0}\right)^{3} {1 \over (1+z)^3} {\left(q_0z + (q_0-1)(\sqrt{1+2q_0z} - 1) \right) \over q^4_0 \sqrt{1+2q_0z}} ~~ dz~, \end{equation} where $zmin$ and $zmax$ are the minimum and maximum redshifts. $zmin = 0.67$, the lower spectral cutoff of the NICMOS G141 grism. The maximum redshift at which an object can be detected, $zmax$, is ${\rm min}(1.9, z2)$, where $z2$ is computed from $D_L(zmax) = D_L(z) \sqrt{f(H\alpha) \over f_{lim}}$, and $D_L$ is the luminosity distance at redshift $z$. $f(\rm H\alpha)$ and $f_{lim}$ are the H$\alpha$ line flux and the flux limit, respectively. $\Omega$ is the solid angle subtended by a single NICMOS camera 3 image. Only the central portion of each grism image samples the full range of wavelengths from $1.1\mu - 1.9\mu$, and the left and right portions of the image sample restricted spectral ranges. The field of view subtended by the central section is $\sim 31^{''}\times 49^{''}$; the two sides of the image subtend roughly $10^{''}\times 49^{''}$ each. Thus the volume must be calculated separately for each portion of the detector. The grism sensitivity is a function of wavelength and this influences the effective volume probed, particularly at the ends of the spectral coverage, where the sensitivity (and hence the available volume) fall off steeply. The estimate of the source density in a luminosity bin of width $\Delta(\log L)$ centered on luminosity $\log L_i$ is simply the sum of the inverse volumes ($\rm 1/V_{max}$) of all the sources with luminosities in the bin. The value of the luminosity function in that bin is \begin{equation} \phi(\log L_i) = {1 \over \Delta (\log L)} \sum_{\mid \log L_j - \log L_i \mid < \Delta(\log L)} { 1 \over V_{max,j}} \end{equation} where index $i$ labels luminosity bins and index $j$ labels galaxies. The variances are computed by summing the squares of the inverse volumes; the error bars of each luminosity bin are the square roots of the variances. The apparent luminosity function must be corrected for incompleteness in the original source catalog. The approach we adopted is similar to that used in Yan et al. (1998). We chose several well-detected emission-line galaxies spanning a range of W$_{\lambda}$ in the form of 2D spectra. We dimmed these template spectra by various factors and added them to the real NICMOS grism images at random locations. We then applied the same detection procedure as in the original analyses (see Paper I) to recover the template spectra. The use of random positions in the simulation allows us to include incompleteness corrections caused by crowding and spatially dependent errors in the sky subtraction and shading corrections. We added the dimmed template spectra to the 2D grism images with a wide range of sensitivities to simulate the true distribution of limiting fluxes in our survey fields. The final averaged detection rate provides a measurement of the dependence of the detection probability on the line flux limits, as well as equivalent width for W$_{rest} > 50$\AA. The incompleteness correction is applied to each bin independently. \begin{figure} \plotone{f1.eps} \caption{H$\alpha$ luminosity function at $ 0.7 < z < 1.9$. The open and filled circles are the data points from our measurements. The open circles represent the raw data and the filled circles are the points corrected for incompleteness. The incompleteness correction is only significant at the faintest luminosity bin. The open triangles show the local H$\alpha$ luminosity function by Gallego et al. (1995). The solid and dashed lines are the best fits to the data at $z \sim 1.3$ and $z \sim 0$ respectively.} \end{figure} In Figure 1 we show our derived luminosity function and the local H$\alpha$ LF as measured by Gallego et al. (1995). Our H$\alpha$ luminosities have been corrected for [NII] contamination using H$\alpha$/[NII]6583,6548 = 2.3 (Kennicutt 1992; Gallego et al. 1996). The solid and dashed lines are the best fits to a Schechter luminosity function at $z \sim 1.3$ and $z \sim 0$, respectively. Gallego et al. (1995) derived $\rm L^\star(H\alpha) = 1.4\times 10^{42}$~erg s$^{-1}$, $\alpha = -1.35$ and $\phi = 6.3\times 10^{-4}$~Mpc$^{-3}$. We obtain $\rm L^\star(H\alpha) = 7\times 10^{42}$~erg s$^{-1}$ and $\phi = 1.7\times 10^{-3}$~Mpc$^{-3}$, assuming a faint end slope, $\alpha$, equal to the local value of $-1.35$. Our sample is not large enough or deep enough to allow an independent determination of $\alpha$. Figure 1 shows strong evolution in the H$\alpha$ luminosity density from $z\sim 0 $ to $z \sim 1.3$. This is no surprise given the evolution in the ultraviolet luminosity density, but our result provides an independent measure of evolution for H$\alpha$ emission alone. The integrated H$\alpha$ luminosity density at $z \sim 1.3$ (our median $z$) is $1.64\times 10^{40}$~h$_{50}$~erg s$^{-1}$ Mpc$^{-3}$, approximately 14 times greater than the local value reported by Gallego et al. (1995). Gronwall (1998) has reported a preliminary measure of the local star formation density derived from the KISS survey that is consistent with the Gallego et al. value. If the faint end slope is as steep as -1.6, as found for the UV luminosity function at $z \sim 3$ (Steidel et al. 1999; Pascarelle, Lanzetta \&\ Fernandez-Soto 1998), the integrated H$\alpha$ luminosity density at $z \sim 1.3$ would be roughly 50\% higher still. Two of the emission-line galaxies in our sample are possible AGN candidates. If we remove the top three objects with the highest H$\alpha$ fluxes from our luminosity function calculation, the integrated H$\alpha$ luminosity density is reduced by $\sim$~40\%, primarily due to the decrease of L$^\ast$(H$\alpha$). \section{Implications for the Evolution of Field Galaxies} We converted the integrated H$\alpha$ luminosity density to a star formation rate (SFR) using the relation from Kennicutt (1999): $\rm {SFR}(M_\odot yr^{-1}) = 7.9 \times 10^{-42} L(H\alpha) (erg~s^{-1}) $. This assumes Case B recombination at $T_e = 10^4$~K and a Salpeter IMF ($0.1 - 100~M_\odot$). This conversion factor is about 10\% smaller than the value listed in Kennicutt (1983), the difference reflecting updated evolutionary tracks. While different choices of stellar tracks introduce modest uncertainties in the conversion of UV and H$\alpha$ luminosities to star formation rates, the choice of different IMFs lead to rather large differences. To make consistent comparisons between our results and those in the literature derived from 1500\AA\ and 2800\AA\ UV continuum luminosity densities, we adopt the relation from Kennicutt (1999): $\rm {SFR}(M_\odot yr^{-1}) = 1.4 \times 10^{-28} L(1500-2800\AA) (erg~s^{-1} Hz^{-1}) $. This relation is appropriate for the Salpeter IMF used to derive the H$\alpha$ conversion factor. \begin{figure} \plotone{f2.eps} \caption{The global volume-averaged star formation rate as a function of redshift without any dust extinction correction. The open squares represent measurements of the 2800\AA\ or 1500\AA\ continuum luminosity density by Lilly et al. (1996), Connolly et al. (1996) and Steidel et al. (1998), whereas the filled squares are the measurements using H$\alpha$~6563\AA\ by Gallego et al. (1995), Tresse \&\ Maddox (1996) and Glazebrook et al. (1998). Our result is shown in the filled circle.} \end{figure} In Figure 2, we plot uncorrected published measurements of the volume-averaged global star formation rate at various epochs. The open squares represent measurements of the 2800\AA\ or 1500\AA\ continuum luminosity density by Lilly et al. (1996), Connolly et al. (1996) and Steidel et al. (1998); the filled squares are the measurements using H$\alpha$~6563\AA\ by Gallego et al. (1995), Tresse \&\ Maddox (1996) and Glazebrook et al. (1998). Our result is shown as a filled circle. The star formation rates shown in Figure 2 are calculated from the luminosity densities integrated over the entire luminosity functions, for both H$\alpha$ and the UV continuum. Lilly et al. and Connolly et al. assumed a faint end slope of $-1.3$ for the UV continuum luminosity functions at $\rm z \le 1$. The 1500\AA\ continuum luminosity function at $z \sim 3 - 4$ measured from Lyman break galaxies by Steidel et al. (1998) has a faint end slope of $-1.6$. Glazebrook et al. (1998) measured H$\alpha$ line fluxes for 13 CFRS galaxies at $z \sim 1$ and concluded that the star formation rate deduced from H$\alpha$ is significantly larger than that derived from the 2800\AA\ continuum luminosity density. Using the conversion factors which we have employed throughout this paper ($7.9\times 10^{-41}$ for H$\alpha$ and $1.4\times 10^{-28}$ for the UV continuum), we estimate that without correcting for extinction the star formation rate derived from their H$\alpha$ luminosity density is a factor of 1.9 higher than that inferred from the 2800\AA\ continuum. The factor of 3 difference in the star formation rate between H$\alpha$ and the UV continuum noted in Glazebrook et al. (1998) includes an extinction correction. As discussed below, the magnitude of the extinction correction is highly dependent on the relative spatial distributions of stars, gas and dust. Pettini et al. (1998) found the star formation rates of Lyman break galaxies at $z \sim 3$ measured from H$\beta$ emission to be $0.7 - 7 \times$ higher than those estimated from the 1500\AA\ continuum. Hogg et al. (1998) and Hammer et al. (1997) derived the [OII]3727 luminosity function and detected strong evolution in the [OII] luminosity density to $z = 1.3$. The conversion factor from the [OII]3727 luminosity to a SFR, however, can differ by an order of magnitude, depending on the metallicity of the gas (Gallagher, Hunter \&\ Bushhouse 1989; Kennicutt 1992). In contrast, H$\alpha$ provides a more direct measure of the star formation by effectively reprocessing the integrated stellar luminosity of galaxies shortward of the Lyman limit. Recent observations at sub-mm wavelengths indicate the star formation rate at $2 < z < 4$ is larger than that inferred from the rest-frame UV luminosity density (Hugh et al. 1998). The contribution of the sub-mm sources to the global star formation history is uncertain at this time, as most of the sub-mm sources do not have secure redshifts (Barger et al. 1999). The clear trend for the longer wavelength determinations of the star formation rate to exceed those based on UV continua is one of the pieces of evidence for significant extinction at intermediate and high redshifts. The amplitude of the extinction correction is quite uncertain (e.g. Heckman et al. 1997; Meuers et al. 1997; Steidel et al. 1998). Our measurement spans $0.7 < z < 1.9$, overlapping with the Connolly et al. photometric redshift sample and allowing a direct comparison between the observed 2800\AA\ luminosity density and that inferred from H$\alpha$. The emission line galaxies selected by our survey have a co-moving number density similar to that of the bright Lyman break galaxies at $z \sim 3$, and a median H magnitude that corresponds to approximately L$^\star$ (Paper I). While we are not comparing the same individual galaxies, the rough correspondence in space density and continuum absolute magnitudes between the UV- and H$\alpha$-selected samples argues that they are drawn from similar or overlapping populations. Our H$\alpha$-based star formation rate is three times larger than the average of the three redshift bins measured by Connolly et al. (1997). The star formation rates derived from line or continuum luminosities depend strongly on the choice of IMF, evolutionary tracks, and stellar atmospheres that are input into a specific spectral evolution model. The relevant issue for the present discussion is the ratio of the star formation rates derived from H$\alpha$ and the 2800\AA\ continuum. As shown by Glazebrook et al. (1998) this ratio differs significantly for the Scalo and Salpeter IMFs and is a function of metallicity. Our choice of the Salpeter IMF comes close to minimizing the difference between the published UV- and our H$\alpha$-derived star formation rates. The use of a Scalo IMF and solar metallicity would increase the apparent discrepancy by a factor of $\sim 2$. The only model considered by Glazebrook et al. that further reduces the H$\alpha$/2800\AA\ star formation ratio is the Salpeter IMF with the Gunn \& Stryker (1983) spectral energy distributions, and this model still leaves us with a factor of $\sim 2$ enhancement in apparent star formation activity measured at H$\alpha$. If we attribute the entire difference to reddening, the total extinction corrections at 2800\AA\ and H$\alpha$ are large and model-dependent. The calculation is sensitive to the relative geometry of the stars, gas and dust, as well as the adopted reddening curve. In the extreme case of a homogeneous foreground screen and a MW or LMC reddening curve, we derive A$_{2800} = 2.1$~magnitudes. In local starburst galaxies, differential extinction between the nebular gas, and stellar continuum, and scattering produce an effective reddening curve that is significantly grayer than the MW or LMC curves (Calzetti, Kinney \&\ Storchi-Bergmann 1994; 1996; Calzetti 1997). The Calzetti reddening law (Calzetti 1997) is appropriate for geometries in which the stars, gas and dust are well mixed. In this model, our estimate of the dust extinction at 2800\AA\ is one to two magnitudes larger than in the simple screen case, and is an uncomfortably large correction compared to results from other methods (e.g., Heckman et al. 1998; Steidel et al. 1999). The properties of the damped Ly$\alpha$ absorbers, diffuse backgrounds and galaxy counts at long wavelengths provide independent constraints on the amount of obscured star formation at large redshifts (e.g. Pei, Fall, \& Hauser 1999; Calzetti \&\ Heckman 1998; Blain et al. 1999). Our measurement of the global star formation rate derived from H$\alpha$ agrees well with the model predictions in Figure 7 of Pei, Fall, \& Hauser. Despite our efforts to quantify the incompleteness of the NICMOS grism sample, some biases remain. The low resolution of the grisms prevent efficient detection of objects with line fluxes above our threshold but with rest-frame W$_{rest} < 50$\AA. The Gallego et al. (1995) survey has a W$_{\lambda}$ threshold of 10\AA. Gallego et al. (1996) find that dwarf amorphous nuclear starbursts have modest equivalent widths but contribute little to the total luminosity density. Some of the compact starburst nuclei will fall below our W$_{\lambda}$ threshold, and these objects can have substantial luminosities. Their compact size mitigates against this somewhat as our spectral resolution is best for point sources. H$\alpha$ spectroscopy of galaxies from the CFRS sample at $z < 0.3$ by Tresse \&\ Maddox (1996) weakly suggests that W$_{\lambda}$(H$\alpha$) and luminosity are correlated. HST imaging and spectroscopic samples all suffer from a bias against low surface brightness objects. The slitless nature of our survey exacerbates the problem as the spectral resolution is a function of apparent source size. The half-light radii of the emission line galaxies in our sample range from 0.2$^{''}$ to 0.7$^{''}$ and the distribution is comparable to that seen in significantly deeper fields, such as the HDF-South and deep NICMOS parallel fields (Yan et al. 1998; Storrie-Lombardi et al. 1999). The principal conclusions of this work are that the H$\alpha$ luminosity density at $z = 1.3$ is an order of magnitude larger than locally, that the global star formation rate derived from our H$\alpha$ measurements exceeds that from the rest-frame UV by a factor of 3 and the implied extinction corrections are substantial. Although the characteristics of this particular data set do not lend themselves to precise comparison between global averaged star formation rates inferred from UV continuum and line-emission, the systematically larger rates inferred from H$\alpha$ at all redshifts point towards significant extinction at rest-frame UV wavelengths. \centerline{\bf 4. Acknowledgments} \vskip 7pt We thank the staff of the Space Telescope Science Institute for their efforts in making this parallel program possible. In particular we thank John Mackenty, Duccio Machetto, Peg Stanley, Doug van Orsow, and the staff of the PRESTO division. We acknowledge useful discussions with M. Fall, J. Gallego, D. Calzetti and R. Marzke. This research was supported, in part, by grants from the Space Telescope Science Institute, GO-7499.01-96A, AR-07972.01-96A and PO423101. HIT acknowledges funding by the Space Telescope Imaging Spectrograph Instrument Definition Team through the National Optical Astronomy Observatories and by the NASA Goddard Space Flight Center.
\section{Introduction} \label{Intro} Perfectly secure communication between two users can be achieved if they share beforehand a common random string of numbers (a key). A big problem in conventional cryptography is the key distribution problem: In classical physics, there is nothing to prevent an eavesdropper from monitoring the key distribution channel passively, without being caught by the legitimate users. Quantum key distribution (QKD) \cite{BB84,Ekert} has been proposed as a new solution to the key distribution problem. In quantum mechanics, there is a well-known ``quantum no-cloning theorem'' which states that it is impossible for anyone (including an eavesdropper) to make a perfect copy of an unknown quantum state \cite{Dieks,WZ}. Therefore, it is generally thought that eavesdropping on a quantum channel will almost surely produce detectable disturbances. \subsection{Prior work on security of QKD} ``The most important question in quantum cryptography is to determine how secure it really is.'' (p. 16 of \cite{BC}) Indeed, there have been many investigations on the issue of security of QKD. Most analyses have dealt with restricted classes of attacks such as single-particle eavesdropping strategies (For a review, see, for example, \cite{Lobook}.), and also the so-called collective attacks \cite{scollect1,scollect2}, where Eve brings each signal particle into interaction with a separate probe, and after hearing the authenticated public discussion between Alice and Bob, measures all the probes together. More recently, the most general type of attacks have been considered. There have been a number of proposed proofs of the {\it unconditional} security of QKD \cite{sdeutsch,LCqkd,smayers1,smayers2,smayers3} based on the laws of quantum mechanics. Note that one should also consider problems of imperfect sources, imperfect measuring devices and noisy channels employed by Alice and Bob. \subsubsection{Why is a proof of security of QKD so difficult?} \label{why} There are many types of eavesdropping strategies. One could imagine that Eve has a quantum computer. In the most general eavesdropping strategy, Eve regards the whole sequence of quantum signals as a single entity. She couples this entity with her probe and then evolves the combined system using a unitary transformation of her choice. Finally, she sends a subsystem to the user(s) and keeps the rest for eavesdropping purposes. Notice that Eve can choose any unitary transformation she likes and yet a secure QKD scheme must defeat all of them. Two major difficulties are expected in a proof of security of QKD. First, Eve tries to evade detection by attributing noises caused by her eavesdropping attack to normal transmission noise. Second, owing to the subtle quantum correlations between Eve and the users, a na\"{\i}ve application of classical arguments may be fallacious. Indeed, there is a well-known paradox---Einstein-Podolsky-Rosen paradox \cite{EPR}---which illustrates clearly the general failing of na\"{\i}ve classical arguments in quantum mechanics. \subsubsection{Two alternative approaches to proving security} \label{ss:two} Roughly speaking, there are two main alternative approaches to proving the unconditional security of QKD. The first approach deals with the most well-known QKD scheme BB84 proposed by Bennett and Brassard \cite{BB84}. The advantage of this approach is that it does not require the employment of quantum computers by Alice and Bob. However, all versions of current proposed proofs of unconditional security based on this approach require the assumption of a perfect photon source \cite{smayers3}. (Earlier versions of \cite{smayers3} have appeared as \cite{smayers1,smayers2} but they are less definite.) Given that a perfect photon source is beyond current technology, proofs based on the first approach (just like those based on the second approach) cannot be directly applied to real-life experiments. See also \cite{Mayao}. The second approach deals with QKD schemes that employ the subtle quantum mechanical correlations---known as ``entanglement''--which have no classical analog. This approach was first suggested in \cite{sdeutsch}, which, however, assumes perfect quantum devices. A more recent paper \cite{LCqkd} addresses this issue of imperfect devices using the idea of fault-tolerant quantum computation and quantum repeaters (i.e., relay stations) \cite{Dur}. It also derives a rigorous bound on Eve's information under the assumption of reliable local quantum computations. Note that the second approach requires Alice and Bob to possess quantum computers, which are well beyond current technology. However, the second approach, as rigorously developed in \cite{LCqkd}, has a number of advantages. First, it extends the range of secure QKD to arbitrarily long distances even with insecure ``quantum repeaters'' (i.e., relay stations). In contrast, such an extension with the first approach will require perfectly secure quantum repeaters. Second, when implemented over a noisy channel, QKD schemes based on the second approach tend to tolerate a larger error rate. Third, a proof of security and the tradeoff between noise and key rate are much easier to work out in the second approach. Fourth, the second approach is conceptually simpler. Finally, some of the techniques developed in the second approach have widespread applications. Indeed, it is plausible that some of those techniques, when properly generalized, can be applied to the first approach. \subsection{Significance of Our results} It has to be said that all previously proposed proofs of security of QKD involve various technical subtleties. Here we present a simple proof of the unconditional security of QKD. The proof, based on the second approach, not only enjoys all the fundamental advantages mentioned above of the recently proposed proof\cite{LCqkd}, but also is conceptually simpler. Besides, our proof gives us an extremely interesting new insight on the well-known ``teleportation'' channel \cite{tele}: With a classical random sampling method, one can assign a set of {\it classical} probabilities to the various error pattern of a {\it quantum} teleportation channel. Besides, the error rate (the probability of having a non-trivial error pattern) for each signal is independent of the identity of the signal being transmitted. This is highly non-trivial because, as noted in subsubsection~\ref{why}, the well-known Einstein-Podolsky-Rosen paradox demonstrates that applications of classical arguments to a quantum problem often lead to fallacies \cite{EPR}. Another potential advantage of our proof is that, with imperfect local quantum computations, it is probable that a longer key can be generated by the current scheme with the same quantum channel. \section{Security requirement and ideas towards a proof} {\bf Definition}: A QKD scheme is said to be unconditionally secure if, for any security parameters $k, l > 0$ chosen by Alice and Bob, they can follow the protocol and construct a verification test such that, for {\it any} eavesdropping attack by Eve that will pass the test with a non-negligible amount of probability, i.e., more than $e^{-k}$, the two following conditions are satisfied: (i) Eve's mutual information with the final key is always negligible, i.e., less than $e^{-l}$ and (ii) the final key is, indeed, essentially random. {\it Remark}: The security parameters $k$ and $l$ depend on how hard Alice and Bob are willing to work towards perfect security (e.g., the size of the messages exchanged between Alice and Bob and the number of rounds of authentication between them) and are, at least in principle, computable from a protocol. \subsection{A simple idea, its problems and our solution} Consider the following simple idea of proof of security of QKD. Alice prepares $r$ quantum signals and encodes their state into a quantum error correcting code (QECC) (see, for example, \cite{BDSW}) of length $n$ which corrects say $t$ errors. In addition, she also prepares $m$ other quantum signals which will be used as test signals. She then {\it randomly} permutes the $N=n+m$ signals and sends them to Bob via a noisy channel controlled by an eavesdropper. Bob publicly announces that he has received all the $N$ signals from Alice. Upon Bob's confirmation of the receipt, Alice publicly announces the location of the $m$ test signals and their specific state. Now, Bob measures the $m$ test signals and computes their error rate, $e_1$. Using the error rate $e_1$, Alice and Bob apply classical random sampling theory in statistics to establish confidence levels for the error rate of the $n$ remaining (i.e., untested) signals and, hence, produce a probabilistic bound on the amount of eavesdropper's information on the encoded $r$ quantum signals. [The point is that, unless there are more than $t$ errors in the QECC, Eve knows absolutely nothing about the encoded state.] If Alice and Bob are satisfied with the degree of security, they measure the $r$ quantum signals to generate an $r$-bit key. This raw idea looks simple, but it is essentially classical. It will work if the following three requirements are satisfied. (1) Each error pattern can be assigned with a classical probability; (2) Error rate of the signals are independent of the actual signals being transmitted (i.e., Eve cannot somehow change a non-trivial error operator to a trivial one depending on which signals are transmitted); (3) The quantum error correction and key generation can be done fault-tolerantly. Since applications of classical arguments could be fallacious, it would be na\"{\i}ve to assign a probability distribution to the set of error patterns without a rigorous mathematical justification. In fact, rather disappointingly, we are unable to establish requirements (1) and (2) for the most general quantum channel. Nonetheless, we manage to complete our proof of security of QKD by the following line of arguments. We notice that requirement (1) has already been established in \cite{LCqkd} for the special case of the transmission of some standard states (halves of so-called EPR pairs). Moreover, it is well-known in quantum information theory that the transmission of any general quantum state can be reduced to that of the standard state and classical communication via a process called {\it teleportation} \cite{tele}, (which will be discussed in subsection~\ref{ss:tele}). Our line of attack is, thus, to establish requirements (1) and (2) for the special case of a teleportation channel only. In other words, we show that, by using teleportation to transmit quantum states through a noisy quantum channel (which may be controlled by an eavesdropper), the error rate [i.e., the probability of having a non-trivial error operator (or Pauli matrix) acting on the transmitted signal, as can be estimated by a classical random sampling procedure] is independent of the quantum state being transmitted. This invariance result ensures that, for a quantum teleportation channel, even an ingenious eavesdropper cannot change its underlying error rate and make it dependent on the identity of the quantum signals being transmitted. This new insight of ours---the ``invariance of the error rate of a quantum teleportation channel''---will be stated as Proposition~5 and discussed in subsequent sections. \subsection{Einstein-Podolsky-Rosen pairs} Readers who are unfamiliar with quantum information should refer to appendix~\ref{physics}. One can measure a quantum bit (or qubit) along any direction and each measurement can give two possible outcomes. An Einstein-Podolsky-Rosen pair of qubits has the following interesting property. If two members of an EPR pair are measured along {\it any} common axis, each member will give a random outcome, and yet, the outcomes of the two members will always be anti-parallel. This is so even when the two members are distantly separated. Such an action at a distance is at the core of the Einstein-Podolsky-Rosen paradox and it defies any simple classical explanation. Now, if two persons, Alice and Bob, share $R$ EPR pairs, they can generate a common random string of number (an $R$-bit key) by measuring each member along some common axis. The laws of quantum mechanics guarantees that, provided that the $R$ pairs are of almost perfect fidelity, the key generated will be almost perfectly random and that Eve will have a negligible amount of information on its value. In fact, we have {\bf Lemma~1}: (Note 28 of \cite{LCqkd}) If Alice and Bob share $R$ EPR pairs of {\it fidelity} at least $ 1 - 2^{-k}$, for a sufficiently large $k$, and they generate an $R$-bit key by measuring these pairs along any common axis, then Eve's mutual information on the final key will be bounded by $2^{-c} + 2^{O(-2k)}$ where $c= k - log_2 \left[ 2R + k + ( 1 / \log_e 2 ) \right]$. {\bf Proof}: In supplementary material of \cite{LCqkd}. So, the Holy Grail of the second approach to secure QKD is to construct a scheme for distributing $R$ almost perfect EPR pairs even in the presence of noises and Eve. \section{Quantum to Classical Reduction Theorem} \subsection{Theory} A proof of security of QKD can be simplified greatly if one can apply well-known powerful techniques in classical probability theory and statistical theory to the problem. However, as noted in subsection \ref{why}, applications of classical arguments to a quantum problem often lead to fallacies. A key ingredient of our current proof is, therefore, a quantum to classical reduction theorem proven in \cite{LCqkd}, which justifies the usage of classical arguments. Let us recapituate this quantum to classical reduction theorem from the viewpoint of ``commuting observables'': Conceptually, classical arguments work because all the observables $ O_i$'s under consideration are diagonal with respect to a {\it single} basis, which we shall call $\cal B$. More concretely, let $M$ be the observable that represents the complete von Neumann measurement along the basis $\cal B$. Since $O_i$'s and $M$ are all diagonal with respect to the basis $\cal B$, they clearly commute with one another. Therefore, the measurement $M$ along basis $\cal B$ will in no way change the outcome of subsequent measurements $O_i$'s. Without loss of generality, we can imagine that such a measurement $M$ is always performed before the measurement of subsequent $O_i$'s. Consequently, the initial state is always a classical mixture of eigenstates of $M$ and, hence, classical arguments carry over directly to a quantum problem. In this sense, the quantum problem has a classical interpretation.\footnote{This quantum to classical reduction theorem is rather subtle. First, the observables $O_i$'s under consideration are coarse-grained observables (i.e. observables with degenerate eigenvalues), rather than fine-grained ones (i.e. observables with non-degenerate eigenvalues). It is {\it a priori} surprising that coarse-graining as a mathematical technique will give a classical interpretation to a quantum problem. Second, the eigenstates of $M$ employed in \cite{LCqkd} are, in fact, the so-called Bell states (see subsection~\ref{application} and appendix~\ref{a:Bell}), which exhibit non-local quantum mechanical correlations. It is {\it a prior} surprising that such a non-local (or quantum mechanical) Bell basis can have a classical interpretation.} Mathematically, this quantum to classical reduction theorem can be stated as the following theorem. {\bf Theorem~2}: \cite{LCqkd} Consider a mixed quantum state described by $\rho$ and a set of one-dimensional non-commuting projection operators $Q_j$ on it. Suppose there exists a complete set of coarse-grained observables $O_i$ of $Q_j$ such that all the $O_i$'s commute with one another. [Here, by coarse-graining, one means that each $O_i$ can be written as a sum of a set of {\it orthogonal} projectors $Q_j$ and by completeness, one means that $\sum_i O_i = I$.] Let us consider a complete von Neumann measurement $M$ which commutes with all $O_i$. [Because of the commutativity of $O_i$'s, such $M$ must exist.] Let $| v_k \rangle$ be the basis vectors of $M$. Then, Theorem~2 says that, for all $i$, we have \begin{equation} {\rm Tr} \left( O_i \rho \right) ={\rm Tr} \left( O_i \sum_k | v_k \rangle \langle v_k| \rho | v_k \rangle \langle v_k| \right). \label{e:theorem2} \end{equation} {\it Remark}: Physically, Theorem~2 says that the probability of all the coarse-grained outcome $O_i$'s are unchanged by a prior complete von Neumann measurement $M$. The full power of Theorem~2, will be demonstrated in Propostion~3. {\bf Proof}: Sketch. By construction, for each $O_i$ there exist a coefficient $\lambda_i$ and a set $K_i$ such that $O_i = \lambda_i \sum_{l \in K_i} | v_l \rangle \langle v_l|$. From the definition of ${\rm Tr} A $ as $\sum_m \langle v_m | A| v_m \rangle$, it is now a simple exercise to establish Eq.~(\ref{e:theorem2}). \subsection{Application to random sampling} \label{application} Consider the following example (example (i) on p. 2054 of \cite{LCqkd}). Suppose two distant observers, Alice and Bob, share a large number, say $N$, pairs of qubits, which may be prepared by Eve. Those pairs may, thus, be entangled with one another in an arbitrary manner and also with the external universe, for example, an ancilla prepared by Eve. How can Alice and Bob estimate the number of singlets in those $N$ pairs? (By the number of singlets, here we mean the expected number of ``yes'' answers if a singlet-or-not measurement were made on each pair individually.) The solution is the following random sampling procedure and proposition. {\it Procedure}: Suppose Alice and Bob randomly pick $m$ of the $N$ pairs and, for each pair, choose randomly one of the three ($x$, $y$ and $z$) axes and measure the two members along it. They publicly announce their outcomes. Let $k$ be the number of anti-parallel outcomes obtained in this random sampling procedure. {\bf Proposition~3}: (in Section VI of supplementary material of \cite{LCqkd}) The fraction of singlets, $f_s$, in the $N$ pairs can be estimated as $(3k -m)/2m$. Furthermore, confidence levels can be deduced from classical statistical theory for a finite population (of $N$ objects). {\bf Proof}: A direct application of Theorem~2. Let us order the $N$ pairs. Consider, for the $i$-th pair, the projection operations $P^i_{\parallel,a}$ and $P^i_{{\rm anti}-\parallel, a}$ for the two coarse-grained outcomes (parallel and anti-parallel) of the measurements on the two members of the pair along the $a$ axis where $a = x, y$ or $z$. A simple but rather important observation is the following: each of these projection operators can be mathematically re-written as linear combination of projection operators along a single basis, namely Bell basis. (See appendix~\ref{a:Bell} for details.) A basis for $N$ ordered pairs of qubits (what we shall call $N$-bell basis) consists of products of Bell basis vectors, each of which is described by a $2N$-bit string. Now, let us consider the operator $M_B$ that represents the action of a complete von Neumann measurement along $N$-Bell basis. Since $M_B$, $P^i_{\parallel,a}$ and $P^i_{{\rm anti}-\parallel, a}$ are diagonal with respect to a single basis ($N$-Bell basis), they clearly commute with each other. Thus, a pre-measurement $M_B$ by Eve along $N$-Bell basis will in no way change the outcome for $P^i_{\parallel,a}$ and $P^i_{{\rm anti}-\parallel, a}$. With any loss of generality, we can assume that such a pre-measurement is always performed before the subsequent measurement of $P^i_{\parallel,a}$ and $P^i_{{\rm anti}-\parallel, a}$. In other words, we have a classical mixture of $N$-Bell basis vectors and classical probability theory refering only to the $N$-Bell basis vectors is, thus, valid. For this reason, estimation of the number of singlets as well as confidence levels of such an estimation can be done by classical statistical theory.~QED. \section{Our secure QKD scheme} We remark that the fraction of singlets, $f_s$, in Proposition~3 has the significance as being the fraction of uncorrupted qubits in a quantum communication channel shared between Alice and Bob in the following situation. Suppose Alice prepares $N$ EPR pairs locally and, afterwards, sends a member of each pair to Bob via a noisy quantum channel controlled Eve. As a result of channel noises and eavesdropping attack, some of the $N$ EPR pairs may be corrupted. Proposition~3 gives us a mathematical estimate of the number of uncorrupted qubits in the actual transmission, based on the random sampling of a small number of transmitted signals. Since quantum error correcting codes (QECCs) exist, it is tempting to construct a secure QKD scheme by, first, using the random sampling procedure to estimate the error rate of the transmission and, second, using a QECC to correct the appropriate number of errors. To ensure that the sampling procedure is, indeed, random, Alice should mix up the test pairs with the pairs in the actual QECC randomly. However, as briefly noted in the Introduction, the above idea implicitly assumes that the following conjecture is true. Let us consider the four error operators $I$, $\sigma_x$, $\sigma_y$ and $\sigma_z$ for each quantum signals transmitted. (See appendix~\ref{physics} for notations.) {\bf Conjecture~4}: The error rate of a quantum communication channel is independent of the signals being transmitted. More precisely, in the current case, one can safely assign a probability for each error pattern in analyzing the security issue of QKD scheme. While such a conjecture is intuitively plausible, we are unaware of any rigorous proof for a general quantum channel. To address this problem, we prove a related but perhaps weaker result concerning a teleportation channel. We make use of the well-known fact that, any quantum signals can always be transmitted to a quantum communication channel via teleportation. \subsection{Teleportation} \label{ss:tele} In teleportation \cite{tele}, a quantum signal is transported via a dual usage of prior ``entanglement'' (i.e. standard EPR pairs shared between the sender, Alice, and the receiver, Bob) and a classical communication channel. The quantum signal in Alice's hand is destroyed by her local measurement, which generates a classical message. This message is then transmitted to Bob via a classical communication channel. Depending on the content of this message, Bob can then re-construct the destroyed quantum signal by applying one of the unitary transformations $I$, $\sigma_x$, $\sigma_y$ and $\sigma_z$ to each of his member of the EPR pairs originally shared with Alice. Two points are noteworthy. First, in teleportation the same prior entanglement is shared by Alice and Bob, independent of the actual quantum signal that will subsequently be transported. Now, since Alice always sends the same standard quantum signal to Bob during the prior sharing part of the teleportation process, the discussion of classical random sampling theory in subsection \ref{application} can be applied directly. Second, the re-construction step in teleportation, if done with reliable quantum computers, will not introduce new errors into the quantum system. Indeed, if Alice and Bob use a {\it noisy} quantum state shared between them for teleportation, for each transmitted signal, the three types of errors $\sigma_x$, $\sigma_y$ and $\sigma_z$ are simply permuted to one another during the re-construction process. This idea is true even for a quantum superposition of error patterns and entanglement with external universe (as specified by the original noisy quantum state shared between them). Let us formulate this result mathematically. Consider the teleportation of a system $\cal S$ consisting of $N$ qubits from Alice to Bob with the most general mixed state $\rho_u$. Without loss of generality, a system decribed by a mixed state can be equivalently described by a pure state of a larger system consisting of the original system and an ancilla. (John Smolin has coined the name ``the Church of the larger Hilbert space'' for this simple but useful observation, which has recently been extensively used \cite{sdeutsch,LCbitcom,Mayersbitcom,Lo}. For instance, the generality of the recent proofs of the impossibility of bit commitment\cite{LCbitcom,Mayersbitcom} and one-out-of-two oblivious transfer\cite{Lo} follows from this idea.) Applying this idea to our current case, the state of original system $\cal S$ (plus an ancilla $\cal R$ with which it is entangled) can be written in the following form (so-called Schmidt decomposition): \begin{equation} | v \rangle_{\cal RS}= \sum_m c_m | w_m \rangle_{\cal R} | v_m \rangle_{\cal S}, \label{v} \end{equation} where $c_m$ are some complex coefficients, $| w_m \rangle_{\cal R}$ and $| v_m \rangle_{\cal S}$ are some basis vectors of the two systems $\cal R$ and $\cal S$ respectively. The initial state $\rho_u$ of the $N$ pairs shared by Alice and Bob can also be {\it purified} in ``the Church of the larger Hilbert space'' as \begin{equation} | u \rangle = \sum_{i_1, i_2 , \cdots, i_N} \sum_j \alpha_{i_1, i_2 , \cdots, i_N,j} | i_1, i_2 , \cdots, i_N \rangle \otimes | j \rangle , \label{two0} \end{equation} where $i_k$ denotes the state of the $k$-th pair and it runs from $\tilde{0} \tilde{0}$ to $\tilde{1}\tilde{1}$, the $| j \rangle$'s form an orthonormal basis for the environment (or an ancilla prepared by Eve), and $\alpha_{i_1, i_2 , \cdots, i_N,j}$ are some complex coefficients. Each state $| u \rangle$ represents a particular mixed state. Note that $| u \rangle $ can be re-written as an entangled sum of a linear superposition of various error patterns. i.e., \begin{equation} | u \rangle = \sum_{i_1, i_2 , \cdots, i_N} \sum_j \alpha_{i_1, i_2 , \cdots, i_N,j} (\prod_k \sigma^{(k)}_{i_k}) | \Psi^- \rangle^N \otimes | j \rangle , \label{u} \end{equation} where $\sigma^{(k)}_{i_k} $ acts on Bob's member of the $k$-th pair as either $I$, $\sigma_x$, $\sigma_y$ or $\sigma_z$ depending on the value of $i_k$, and $| \Psi^- \rangle$ denotes an EPR pair. With such notations, one can prove our main proposition. {\bf Proposition~5: Invariance of error rate under teleportation.} In the above notations, suppose the system $\cal S$ (described by $| v \rangle_{\cal RS}= \sum_m c_m | w_m \rangle_{\cal R} | v_m \rangle_{\cal S}$ of the combined system $\cal R$ and $\cal S$ in Eq.~(\ref{v})) is teleported using the $N$ pairs shared by Alice and Bob (described by $| u \rangle $ of the combined system of the $N$ pairs and Eve's ancilla in Eq.~(\ref{u})). Suppose further that the classical outcome of Alice's measurements is $\{j_k\}$. i.e., she informs Bob to use the operator $ \prod_k \sigma^{(k)}_{j_k}$ for the re-contruction process. Then, Bob's re-constructed state for the combined system $\cal R$, $\cal S $ and $\cal E$ can be described by \begin{equation} \sum_m c_m | w_m \rangle_{\cal R} \sum_{i_1, i_2 , \cdots, i_N} \sum_j \alpha_{i_1, i_2 , \cdots, i_N,j} \left[ \prod_k \left( \sigma^{(k)}_{j_k}\sigma^{(k)}_{i_k}\sigma^{(k)}_{j_k} \right) \right] | v_m \rangle_{\cal S} \otimes | j \rangle. \end{equation} {\it Remark}: The set of complex coefficients $c_m \alpha_{i_1, i_2 , \cdots, i_N,j} $ remain totally unchanged under teleportation. For each teleportation outcome labelled by $\{j_k\}$, the only real change lies in the conjugation action in the error operator acting on the subsystem $\cal S$. i.e., $\sigma^{(k)}_{i_k} \to \sigma^{(k)}_{j_k}\sigma^{(k)}_{i_k}\sigma^{(k)}_{j_k}$ for each $k$. (Recall that $\sigma^{(k)}_{j_k}$ is always its own inverse.) Since under such conjugation the trivial error operator (i.e., the identity $I$) is invariant and the three non-trivial error operators $\sigma_x$, $\sigma_y$ and $\sigma_z$ are permuted to one another, the error rate of the teleported signal is exactly the same as the original $N$ EPR pairs. {\bf Proof of Proposition~5}: A straightforward exercise in quantum information theory \cite{tele}, which we will skip here. \subsection{Procedure of our secure QKD scheme} Having established Proposition~5, we now present the procedure of our secure QKD scheme. 1) Alice prepares $N$ EPR pairs and sends a member of each pair to Bob through a noisy channel. [In theory, quantum repeaters \cite{Dur} and two-way schemes for so-called entanglement purification \cite{BB84} (a generalization of quantum error correcting codes) could be used in this step. The error rate here can, therefore, be made to be very small and the scheme works even for arbitrarily long distances.) 2) Bob publicly announces his receipt of the $N$ quantum signals. 3) Alice {\it randomly} picks $m$ of the $N$ EPR pairs for testing. She publicly announces her choice to Bob. For each pair, Alice and Bob randomly pick one of the three ($x$, $y$, and $z$) axes and perform a measurement on the two members along it. 4) Alice and Bob publicly announce their measurement outcomes and use classical sampling theory to estimate the error rate in the transmission. {\it Remark}: Proposition~3 allows Alice and Bob to apply classical sampling theory to the quantum problem at hand to estimate the error rate of the untested particles. Alice and Bob then proceed with quantum error correction in the next step. 5) Alice prepares say $R$ EPR pairs and encodes the $R$ halves of the pairs (i.e., one member from each pair) by a quantum error correcting code (QECC) into $N-m$ qubits. {\it Remark}: The requirement of QECC will be discussed in subsection~\ref{ss:ftqc}. 6) Alice teleports the $N-m$ qubits to Bob via the remaining $N-m$ pairs that they share. {\it Remark}: Proposition~5 guarantees the invariance of error rate under teleportation. So, the estimate done by Alice and Bob in step 4) remains valid. 7) Alice and Bob perform fault-tolerant quantum computation to generate a random $R$-bit key by measuring the state of the $R$ encoded EPR pairs along a prescribed common axis (say the $z$ axis). \subsection{Fault-tolerant quantum computation} \label{ss:ftqc} From Proposition~3 and~5, it is quite clear that, assuming reliable local quantum computers, our scheme works perfectly. However, since local quantum computations may be imperfect, errors may be generated during the teleportation and key generation, i.e., steps 6) and 7). One can easily take those local errors into account by a choice of QECC with generous error-correcting and fault-tolerant capabilities. The point is that we have a very specific and short computation in mind (measurement along $z$ axis only and no unitary computation at all). Based on any realistic error model for quantum computers and concrete choice of QECC, one can give a generous upper bound on the number of local errors due to imperfect quantum computation. With a fault-tolerant implementation, the total number of errors in the whole process (transmission, teleportation and key generation) can be bounded. Therefore, provided that our QECC has a sufficiently generous error-correcting and fault-tolerant capabilities, security is guaranteed. [To be precise, in step 5), the $R$ EPR pairs should be prepared fault-tolerantly in an {\it encoded} form rather than in an unencoded form.] We remark that, since the required quantum computation here is much simpler than in \cite{LCqkd}, the present QKD scheme may be more efficient than the one there. \section{Concluding Remarks} In summary, we have presented a simple proof of the unconditional security of quantum key distribution, i.e., ultimate security against the most general eavesdropping attack and the most general types of noises. Our scheme allows secure QKD over arbitrarily long distances, but it requires Alice and Bob to have reliable quantum computers, which is far beyond current technology. However, to put things in perspective, all proposed proofs of security of QKD involve assumptions (such as ideal sources) that are beyond current technologies. Notice that some of the techniques developed here and in \cite{LCqkd} have widespread applications. For example, Note 21 of \cite{LCqkd} shows that teleportation is a powerful technique against the quantum Trojan Horse attack. A new application---use random sampling and random teleportation to prove the feasibility of a general two-party fault-tolerant quantum computation even in the presence of eavesdroppers---will be discussed in appendix~\ref{a:two-party}. In fact, some of the results are applicable even to the case when Alice and Bob do not have a quantum computer. A good example is a quantitative statement on the tradeoff between information gain and disturbance in BB84 \cite{LCqkd}. We particularly thank P. W. Shor for inspiring discussions. Very helpful comments from C. H. Bennett, H. F. Chau, and John Smolin are also gratefully acknowledged.
\section{INTRODUCTION} This is the seventh part of our eight presentations in which we consider applications of methods from wavelet analysis to nonlinear accelerator physics problems. This is a continuation of our results from [1]-[8], in which we considered the applications of a number of analytical methods from nonlinear (local) Fourier analysis, or wavelet analysis, to nonlinear accelerator physics problems both general and with additional structures (Hamiltonian, symplectic or quasicomplex), chaotic, quasiclassical, quantum. Wavelet analysis is a relatively novel set of mathematical methods, which gives us a possibility to work with well-localized bases in functional spaces and with the general type of operators (differential, integral, pseudodifferential) in such bases. In contrast with parts 1--4 in parts 5--8 we try to take into account before using power analytical approaches underlying algebraical, geometrical, topological structures related to kinematical, dynamical and hidden symmetry of physical problems. We described a number of concrete problems in parts 1--4. The most interesting case is the dynamics of spin-orbital motion (part 4). In section 2 we consider dynamical consequences of covariance properties regarding to relativity (kinematical) groups and continuous wavelet transform (CWT) (in section 3) as a method for the solution of dynamical problems. We introduce the semidirect product structure, which allows us to consider from general point of view all relativity groups such as Euclidean, Galilei, Poincare. Then we consider the Lie-Poisson equations and obtain the manifestation of semiproduct structure of (kinematic) symmetry group on dynamical level. So, correct description of dynamics is a consequence of correct understanding of real symmetry of the concrete problem. We consider the Lagrangian theory related to semiproduct structure and explicit form of variation principle and corresponding (semidirect) Euler-Poincare equations. In section 3 we consider CWT and the corresponding analytical technique which allows to consider covariant wavelet analysis. In part 8 we consider in the particular case of affine Galilei group with the semiproduct structure the corresponding orbit technique for constructing different types of invariant wavelet bases. \section{Dynamics on Semidirect Products} Relativity groups such as Euclidean, Galilei or Poincare groups are the particular cases of semidirect product construction, which is very useful and simple general construction in the group theory [9]. We may consider as a basic example the Euclidean group $SE(3)=SO(3)\bowtie{\bf R}^3$, the semidirect product of rotations and translations. In general case we have $S=G\bowtie V$, where group G (Lie group or automorphisms group) acts on a vector space V and on its dual $V^*$. Let $V$ be a vector space and $G$ is the Lie group, which acts on the left by linear maps on V (G also acts on the left on its dual space $V^*$). The Lie algebra of S is the semidirect product Lie algebra, $s={\mathcal{G}} \bowtie V$ with brackets $ [(\xi_1,v_1),(\xi_2,v_2)]=([\xi_1,\xi_2],\xi_1v_2-\xi_2v_1), $ where the induced action of $\mathcal{G}$ by concatenation is denoted as $\xi_1 v_2$. Let $(g,v)\in S=G\times V, \quad (\xi,u)\in s={\mathcal{G}}\times V$, $(\mu,a)\in s^*={\mathcal G}^*\times V^*$, $g\xi=Ad_g\xi$, $g\mu=Ad^*_{g^{-1}}\mu$, $ga$ denote the induced left action of $g$ on $a$ (the left action of G on V induces a left action on $V^*$ --- the inverse of the transpose of the action on V), $\rho_v: {\cal G}\to V$ is a linear map given by $\rho_v(\xi)=\xi v$, $\rho^*_v: V^*\to{\cal G}^*$ is its dual. Then adjoint and coadjoint actions are given by simple concatenation: $(g,v)(\xi,u)=(g\xi,gu-(g\xi)v)$, $(g,v)(\mu,a)=(g\mu+\rho^*_v(ga),ga)$. Also, let be $\rho^*_v a=v\diamond a\in{\cal G^*}$ for $a\in V^*$, which is a bilinear operation in $v$ and $a$. So, we have the coadjoint action: $ (g,v)(\mu,a)=(g\mu+v\diamond(ga),ga). $ Using concatenation notation for Lie algebra actions we have alternative definition of $v\diamond a\in{\mathcal G}^*$. For all $v\in V$, $a\in V^*$, $\eta\in{\mathcal G}$ we have $ <\eta a,v>=-<v\diamond a, \eta> $. Now we consider the manifestation of semiproduct structure of symmetry group on dynamical level. Let $F,G$ be real valued functions on the dual space ${\mathcal G}^*$, $\mu\in{\mathcal G}^*$. Functional derivative of F at $\mu$ is the unique element $\delta F/\delta\mu\in{\mathcal G}$: $ \lim_{\epsilon\to 0} \lbrack F(\mu+\epsilon\delta\mu)-F(\mu)\rbrack/\epsilon= <\delta\mu,{\delta F}/{\delta\mu}> $ for all $\delta\mu\in{\mathcal G}^*$, $<,>$ is pairing between $\mathcal G^*$ and $\mathcal G$. Define the $(\pm)$ Lie-Poisson brackets by $\{F,G\}_\pm(\mu)=\pm <\mu,\lbrack{\delta F}/{\delta\mu}, {\delta G}/{\delta\mu}\rbrack>.$ The Lie-Poisson equations, determined by $\dot{F}=\{F,H\}$ or intrinsically $\dot{\mu}=\mp ad^*_{\partial H/\partial\mu}\mu.$ For the left representation of G on V $\pm$ Lie-Poisson bracket of two functions $f,k: s^*\to {\bf R}$ is given by \begin{eqnarray} &&\{f,k\}_{\pm}(\mu, a)=\pm <\mu,\lbrack\frac{\delta f}{\delta\mu}, \frac{\delta k}{\delta\mu}\rbrack>\\ &&\pm <a,\frac{\delta f}{\delta\mu}\frac{\delta k}{\delta a}- \frac{\delta k}{\delta\mu}\frac{\delta f}{\delta a}>,\nonumber \end{eqnarray} where $\delta f/\delta\mu\in{\mathcal G}$, $\delta f/\delta a\in V$ are the functional derivatives of f. The Hamiltonian vector field of $h: s^*\in{\bf R}$ has the expression $X_h(\mu,a)=\mp(ad^*_{\delta h/\delta\mu}\mu-{\delta h}/{\delta a}\diamond a, -{\delta h}/{\delta\mu}a).$ Thus, Hamiltonian equations on the dual of a semidirect product are [9]: \begin{eqnarray}\label{eq:mua} \dot{\mu}=\mp ad^*_{\delta h / \delta\mu}\mu\pm\frac{\delta h}{\delta a}\diamond a,\quad \dot{a}=\pm\frac{\delta h}{\delta\mu} a \end{eqnarray} So, we can see the explicit contribution to the Poisson brackets and the equations of motion which come from the semiproduct structure. Now we consider according to [9] Lagrangian side of a theory. This approach is based on variational principles with symmetry and is not dependent on Hamiltonian formulation, although it is demonstrated in [9] that this purely Lagrangian formulation is equivalent to the Hamiltonian formulation on duals of semidirect product (the corresponding Legendre transformation is a diffeomorphism). We consider the case of the left representation and the left invariant Lagrangians ($\ell$ and L), which depend in additional on another parameter $a\in V^*$ (dynamical parameter), where V is representation space for the Lie group G and L has an invariance property related to both arguments. It should be noted that the resulting equations of motion, the Euler-Poincare equations, are not the Euler-Poincare equations for the semidirect product Lie algebra ${\mathcal G}\bowtie V^*$ or ${\mathcal G}\bowtie V$. So, we have the following: {\bf 1.} There is a left representation of Lie group G on the vector space V and G acts in the natural way on the left on $TG\times V^*: h(v_g,a)=(hv_g,ha)$. {\bf 2.} The function $L: TG\times V^*\in{\bf R}$ is the left G-invariant. {\bf 3.} Let $a_0\in V^*$, Lagrangian $L_{a_0}: TG\to{\bf R}$, $L_{a_0}(v_g)=L(v_g,a_0)$. $L_{a_0}$ is left invariant under the lift to TG of the left action of $G_{a_0}$ on G, where $G_{a_0}$ is the isotropy group of $a_0$. {\bf 4.} Left G-invariance of L permits us to define $\ell:{\mathcal G}\times V^*\to{\bf R}$ by $\ell(g^{-1}v_g,g^{-1}a_0)=L(v_g,a_0).$ This relation defines for any $\ell:{\mathcal G}\times V^*\to{\bf R}$ the left G-invariant function $L: TG\times V^*\to{\bf R}$. {\bf 5.} For a curve $g(t)\in G$ let be $\xi(t):=g(t)^{-1}\dot{g}(t)$ and define the curve $a(t)$ as the unique solution of the following linear differential equation with time dependent coefficients $\dot{a}(t)=-\xi(t)a(t),$ with initial condition $a(0)=a_0$. The solution can be written as $a(t)=g(t)^{-1}a_0$. Then we have four equivalent descriptions of the corresponding dynamics: {\bf 1.} If $a_0$ is fixed then Hamilton's variational principle $\delta\int_{t_1}^{t_2}L_{a_0}(g(t),\dot{g}(t){\rm d}t=0$ holds for variations $\delta g(t)$ of $g(t)$ vanishing at the endpoints. {\bf 2.} $g(t)$ satisfies the Euler-Lagrange equations for $L_{a_0}$ on G. {\bf 3.} The constrained variational principle $\delta\int_{t_1}^{t_2}\ell(\xi(t),a(t)){\rm d}t=0$ holds on ${\mathcal G}\times V^*$, using variations of $\xi$ and $a$ of the form $\delta\xi=\dot{\eta}+[\xi,\eta]$, $\delta a=-\eta a$, where $\eta(t)\in{\mathcal G}$ vanishes at the endpoints. {\bf 4.} The Euler-Poincare equations hold on ${\mathcal G}\times V^*$ \begin{equation} \frac{{\rm d}}{{\rm d}t}\frac{\delta\ell}{\delta\xi}= ad_\xi^*\frac{\delta\ell}{\delta\xi}+\frac{\delta\ell}{\delta a}\diamond a \end{equation} So, we may apply our wavelet methods either on the level of variational formulation or on the level of Euler-Poincare equations. \section{Continuous Wavelet Transform} Now we need take into account the Hamiltonian or Lagrangian structures related with systems (2) or (3). Therefore, we need to consider generalized wavelets, which allow us to consider the corresponding structures instead of compactly supported wavelet representation from parts 1-4. In wavelet analysis the following three concepts are used now: 1).\ a square integrable representation $U$ of a group $G$, 2).\ coherent states (CS) over G, 3).\ the wavelet transform associated to U. We consider now their unification [10]-[12]. Let $G$ be a locally compact group and $U_a$ strongly continuous, irreducible, unitary representation of G on Hilbert space ${\mathcal H}$. Let $H$ be a closed subgroup of $G$, $X=G/H$ with (quasi) invariant measure $\nu$ and $\sigma: X=G/H\to G$ is a Borel section in a principal bundle $G\to G/H$. Then we say that $U$ is square integrable $mod(H,\sigma)$ if there exists a non-zero vector $\eta\in{\mathcal H}$ such that $0<\int_X|<U(\sigma(x))\eta|\Phi>|^2{\rm d}\nu(x)=<\Phi|A_\sigma\Phi>\ <\infty,$ $ \forall\Phi\in{\mathcal H}$. Given such a vector $\eta\in{\mathcal H}$ called admissible for $(U,\sigma)$ we define the family of (covariant) coherent states or wavelets, indexed by points $x\in X$, as the orbit of $\eta$ under $G$, though the representation $U$ and the section $\sigma$ [10]-[12]: $S_\sigma={\eta_{\sigma(x)}=U(\sigma(x))\eta|x\in X}$. So, coherent states or wavelets are simply the elements of the orbit under U of a fixed vector $\eta$ in representation space. We have the following fundamental properties: {\bf 1.}Overcompleteness: the set $S_\sigma$ is total in ${\mathcal H}:(S_\sigma)^\perp={0}$. {\bf 2.} Resolution property: the square integrability condition may be represented as a resolution relation: $\int_X|\eta_\sigma(x)><\eta_{\sigma(x)}|{\rm d}\nu(x)=A_\sigma,$ where $A_\sigma$ is a bounded, positive operator with a densely defined inverse. Define the linear map $W_\eta: {\mathcal H}\to L^2(X,{\rm d}\nu),(W_\eta\Phi)(x)=<\eta_{\sigma(x)}|\Phi>.$ Then the range $H_\eta$ of $W_\eta$ is complete with respect to the scalar product $<\Phi|\Psi>_\eta=<\Phi|W_\eta A^{-1}_\sigma W^{-1}_\eta\Psi>$ and $W_\eta$ is unitary operator from ${\mathcal H}$ onto ${\mathcal H}_\eta$. $W_\eta$ is Continuous Wavelet Transform (CWT). {\bf 3.} Reproducing kernel. The orthogonal projection from $L^2(X,{\rm d}\nu)$ onto ${\mathcal H}_\eta$ is an integral operator $K_\sigma$ and $H_\eta$ is a reproducing kernel Hilbert space of functions: $\Phi(x)=\int_XK_\sigma(x,y)\Phi(y){\rm d}\nu(y), \quad \forall \Phi\in{\mathcal H}_\eta.$ The kernel is given explicitly by $K_\sigma(x,y)=<\eta_{\sigma(x)}A_\sigma^{-1}\eta_{\sigma(y)}>$, if $\eta_{\sigma(y)}\in D(A^{-1}_\sigma)$, $\forall y\in X$. So, the function $\Phi\in L^2(X,{\rm d}\nu)$ is a wavelet transform (WT) iff it satisfies this reproducing relation. {\bf 4.} Reconstruction formula. The WT $W_\eta$ may be inverted on its range by the adjoint operator, $W_\eta^{-1}=W_\eta^*$ on ${\mathcal H}_\eta$ to obtain for $\eta_{\sigma(x)}\in D(A_\sigma^{-1})$, $\forall x\in X$ $W_\eta^{-1}\Phi=\int_X\Phi(x)A_\sigma^{-1}\eta_{\sigma(x)}{\rm d}\nu(x), \ \Phi\in{\mathcal H}_\eta.$ This is inverse WT. If $A_\sigma^{-1}$ is bounded then $S_\sigma$ is called a frame, if $A_\sigma=\lambda I$ then $S_\sigma$ is called a tight frame. This two cases are generalization of a simple case, when $S_\sigma$ is an (ortho)basis. The most simple cases of this construction are:\\ {\bf 1.} $H=\{e\}$. This is the standard construction of WT over a locally compact group. It should be noted that the square integrability of U is equivalent to U belonging to the discrete series. The most simple example is related to the affine $(ax+b)$ group and yields the usual one-dimensional wavelet analysis $[\pi(b,a)f](x)=\frac{1}{\sqrt{a}}f\left(\frac{x-b}{a}\right).$ For $G=SIM(2)={\bf R}^2\bowtie({\bf R}^{+}_*\times SO(2))$, the similitude group of the plane, we have the corresponding two-dimensional wavelets. {\bf 2.} $H=H_\eta$, the isotropy (up to a phase) subgroup of $\eta$: this is the case of the Gilmore-Perelomov CS. Some cases of group G are: {\bf a).} Semisimple groups, such as SU(N), SU(N$|$M), SU(p,q), Sp(N,{\bf R}). {\bf b).} the Weyl-Heisenberg group $G_{WH}$ which leads to the Gabor functions, i.e. canonical (oscillator)coherent states associated with windowed Fourier transform or Gabor transform (see also part 6): $[\pi(q,p,\varphi)f](x)=\exp(i\mu(\varphi-p(x-q))f(x-q)$. In this case H is the center of $G_{WH}$. In both cases time-frequency plane corresponds to the phase space of group representation. {\bf c).} The similitude group SIM(n) of ${\bf R}^n(n\ge3)$: for $H=SO(n-1)$ we have the axisymmetric n-dimensional wavelets. {\bf d).} Also we have the case of bigger group, containing both affine and We\-yl-\-Hei\-sen\-berg group, which interpolate between affine wavelet analysis and windowed Fourier analysis: affine Weyl--Heisenberg group [12]. {\bf e).} Relativity groups. In a nonrelativistic setup, the natural kinematical group is the (extended) Galilei group. Also we may adds independent space and time dilations and obtain affine Galilei group. If we restrict the dilations by the relation $a_0=a^2$, where $a_0, a$ are the time and space dilation we obtain the Galilei-Schr\"odinger group, invariance group of both Schr\"odinger and heat equations. We consider these examples in the next section. In the same way we may consider as kinematical group the Poincare group. When $a_0=a$ we have affine Poincare or Weyl-Poincare group. Some useful generalization of that affinization construction we consider for the case of hidden metaplectic structure in part 6. But the usual representation is not square--integrable and must be modified: restriction of the representation to a suitable quotient space of the group (the associated phase space in our case) restores square -- integrability: $G\longrightarrow$ homogeneous space. Our goal is applications of these results to problems of Hamiltonian dynamics and as consequence we need to take into account symplectic nature of our dynamical problem. Also, the symplectic and wavelet structures must be consistent (this must be resemble the symplectic or Lie-Poisson integrator theory). We use the point of view of geometric quantization theory (orbit method) instead of harmonic analysis. Because of this we can consider (a) -- (e) analogously. In next part we consider construction of invariant bases. We are very grateful to M.~Cornacchia (SLAC), W.~Her\-r\-man\-nsfeldt (SLAC) Mrs. J.~Kono (LBL) and M.~Laraneta (UCLA) for their permanent encouragement.
\section{Introduction} The most important difference between mm and cm interferometry is the effect of the troposphere at mm wavelengths. The optical depth of the troposphere becomes significant below 1 cm, leading to increased system temperatures due to atmospheric emission, and increased demands on gain calibration due to variable opacity (Yun et al., 1998, Kutner and Ulich 1981). Even more dramatic is the effect of the troposphere on interferometer phases. Variations in the tropospheric water vapor column density lead to variations in electronic pathlength, and hence variations in interferometric phase. This can cause loss of amplitude of the cross correlations, or `visibilities', over the integration time (`coherence'), reduced spatial resolution (`seeing'), and pointing errors (`anomalous refraction'). Until recently, mm interferometric arrays have been restricted to maximum baselines of a few hundred meters. Diffraction limited resolution images at mm wavelengths can be made on such baselines, since the tropospheric phase fluctuations on such baselines at good observatory sites are typically one radian or less under good weather conditions. However, many existing mm arrays are expanding to baselines of 1 km or more, and the planned large millimeter array facility at the Chajnantor site in Chile will have baselines as long at 10 km. Even at this premium site, phase fluctuations due to the troposphere will be larger than 1 radian at 230 GHz on these baselines, except under the best weather conditions. Hence, tropospheric seeing will preclude diffraction limited resolution imaging at frequencies of 230 GHz for arrays larger than about 1 km, even at the best possible sites, if no corrections are made for tropospheric phase noise. At higher frequencies, the maximum baselines which would permit uncorrected observations would be even shorter. In this paper we review the theory of tropospheric phase noise in mm interferometry, along with examples showing tropospheric induced phase fluctuations and their effect on images made with interferometric arrays. We then consider three techniques for reducing tropospheric phase noise: (i) Fast Switching phase calibration, (ii) Paired Array phase calibration, and (iii) radiometric phase calibration. The first two techniques entail using celestial calibration sources near the target source, with either a calibration cycle time fast enough to `stop' the troposphere (Fast Switching), or by using some of the antennas as a tropospheric `calibration array' (Paired Array). We present extensive observational data using the Very Large Array (VLA) in Socorro, NM, USA, at 22 GHz and 43 GHz designed to test the efficacy of these first two techniques on baselines longer than a few km. The radiometric phase correction technique entails real-time estimation of the precipitable water vapor content along each antenna's line of sight through the troposphere via a radiometric measurement of the brightness temperature of the atmosphere above each antenna. A number of issues are addressed, including: (i) the required radiometric sensitivity as a function of frequency and site quality, (ii) the constraints on ancillary data, such as atmospheric data and models, in order to perform an absolute radiometric phase correction, and (iii) the limitations to making radiometric phase corrections by calibrating the relationship between brightness temperature fluctuations and interferometric phase using celestial sources. \section{A General Description of the Troposphere and the Mean Tropospheric Effect on Interferometric Phase} The troposphere is the lowest layer of the atmosphere, extending from the ground to the stratosphere at an elevation of 7 km to 10 km. The temperature decreases with altitude in this layer, clouds form, and convection can be significant (Garratt 1992). The troposphere is composed predominantly of N$_2$, O$_2$, plus trace gases such as water vapor, N$_2$O, and CO$_2$, plus particulates such as liquid water and dust in clouds. The troposphere becomes increasingly opaque with increasing frequency, mostly due to absorption by O$_2$ and H$_2$O. Figure 1 shows models of the atmospheric transmission at cm and mm wavelengths for the VLA site at 2150 m altitude, and the planned millimeter array (MMA) site in Chile at 4600 m altitude (Holdaway and Pardo 1997, Liebe 1989). The plot shows a series of strong absorption lines including the water lines at 22 GHz and 183 GHz, and the O$_2$ lines at 60 GHz and 118 GHz, plus a systematic decrease in the transmission with increasing frequency between the lines. This `pseudo-continuum' opacity is due to the sum of the pressure broadened line wings of a multitude of sub-mm and IR lines of water vapor. The plot for the MMA site at Chajnantor in Chile includes the typical value for the column density of precipitable water vapor, w$_{\circ}$ = 1 mm, while the water vapor column for the VLA site is assumed to be w$_{\circ}$ = 4 mm, where precipitable water vapor (PWV) = the depth of the water vapor if converted to the liquid phase. Figure 2 shows the relative contributions from water vapor and dry air (O$_2$ plus other trace gases) for the VLA site. Below 130 GHz, both O$_2$ and H$_2$O contribute significantly to the optical depth. Above 130 GHz, H$_2$O dominates the optical depth. The troposphere has a non-unit refractive index, $n$. The refractive index is defined by the phase change experienced by an electromagnetic wave, $\phi_e$, propagating over a physical distance, D: $$ \phi_e = {{2 \pi}\over{\lambda}} \times n \times \rm D, $$ or, in terms of `electrical pathlength', L$_{\rm e}$: $$ \rm L_{e} = \lambda \times {{\phi_e}\over{2 \pi}} = {\it n} \times D $$ The refractive index of air is non-dispersive (ie, independent of frequency) except near the strong resonant water and O$_2$ lines, and is typically given as a difference with respect to vacuum ($n_{vacuum}$ $\equiv$ 1), in parts per million, $N$, as (Waters 1976): $$ N \equiv (n - 1) \times 10^{6} $$ The index of refraction of air is typically separated into the dry air component, $N_{d}$, and the water vapor component, $N_{wv}$. These terms behave as (Waters 1976, Bean and Dutton 1968): \begin{eqnarray} N_{d} = 2.2 \times 10^{5} \times \rho_{tot} \nonumber \\ N_{wv} = 1.7 \times 10^{9} \times {{\rho_{wv}}\over{\rm T_{\rm atm}}} \nonumber \end{eqnarray} where $\rho_{tot}$ is the total mass density in gm cm$^{-3}$, and $\rho_{wv}$ is the water vapor mass density. The inverse dependence on temperature for $N_{wv}$ is due to the increased effect of collisions on the (mis-)alignment of the permanent electric dipole moments of the water molecules with increasing temperature (Waters 1976, Bean and Dutton 1968). A detailed derivation of these relationships can be found in Thompson, Moran, and Swenson (1986). For water vapor alone, it can be shown that $\rho_{wv}$ = ${\rm w}\over{\rm D}$. Using the equations above then leads to the relationship between the electrical pathlength, L$_{\rm e}$, and the precipitable water vapor column, w: \begin{eqnarray} \rm L_e = 1.7\times10^3 {{w}\over{T_{\rm atm}}} \approx 6.3\times\rm w \nonumber \end{eqnarray} or: \begin{equation} \phi_e \approx {{12.6 \pi}\over{\lambda}} \times \rm w \end{equation} for T$_{\rm atm}$ $\approx$ 270 K. This relation between electrical pathlength and precipitable water vapor column has been verified experimentally for a number of atmospheric conditions (Hogg, Guiraud, and Decker 1981). \section{Phase Variations due to the Troposphere} Variations in precipitable water vapor lead to variations in the effective electrical path length, corresponding to variations in the phase of an electromagnetic wave propagating through the troposphere (Tatarskii 1978). Such variations are seen as `phase noise' by radio interferometers. Since the troposphere is non-dispersive, the phase contribution by a given amount of water vapor increases linearly with frequency (except in the vicinity of the strong water lines). Hence, tropospheric phase variations are most prominent for mm and sub-mm interferometers, and can be the limiting factor for the coherence time and spatial resolution of mm interferometers (Hinder and Ryle 1971, Lay 1997, Wright 1996). The standard model for tropospheric phase fluctuations involves variations in the water vapor column density in a turbulent layer in the troposphere with a mean height, h$_{\rm turb}$, and a vertical extent, W, which moves at some velocity, v$_a$. This model includes the `Taylor hypothesis', or `frozen screen approximation', which states that: `if the turbulent intensity is low and the turbulence is approximately stationary and homogeneous, then the turbulent field is unchanged over the atmospheric boundary layer time scales of interest and advected with the mean wind' (Taylor 1938, Garratt 1992). Under this assumption one can relate temporal and spatial phase fluctuations with a simple Eulerian transformation between baseline length, b, and v$_a$:~ b = v$_a$$\times$time. In the following sections we adopt a value of v$_a$ = 10 m s$^{-1}$. This process is shown schematically in Figure 3. A demonstration of tropospheric phase fluctuations is shown in Figure 4 for observations made with the VLA at 22 GHz. The VLA is an aperture synthesis array comprised of 27 parabolic antennas of 25m diameter, operating at frequencies from 75 MHz to 50 GHz (Napier, Thompson, and Ekers 1983). The antennas are situated in a `Y' pattern, along three arms situated north, southwest, and southeast. The maximum physical baseline in the largest configuration is 33 km. The complex cross correlations of the electric fields measured at each antenna are calculated between all pairs of antennas (`interferometers') as a function of time. The Fourier transform of these measurements of the spatial coherence function of the electric field then gives the sky brightness distribution (Clark 1998). For these observations, two subarrays were employed, one observing the celestial calibrator 0423+418, and the second observing the calibrator 0432+416. The sub-arrays were `inter-laced', meaning that every second antenna along each arm of the array observed a given source. Antenna-based phase and amplitude solutions were derived from the data in each sub-array using self-calibration with an averaging time of 30 sec (Cornwell 1998). Figure 4 shows the antenna-based phase solutions from two pairs of neighboring antennas along the southwest arm. The antennas at stations W16 and W4 were observing 0423+418 while the antennas at W18 and W6 were observing 0432+416. For adjacent pairs of antennas (W16-W18 and W6-W4) the temporal variations in the phase track each other closely. This close relationship for phase variations between neighboring antennas in the two different subarrays is the signature that the phase variations are primarily tropospheric in origin, and are correlated on relevant timescales and baseline lengths. An important aspect of tropospheric phase fluctuations arising from the Taylor hypothesis is the relationship between the amplitude of the fluctuations and the time scale: large amplitude fluctuations occur over long periods and are partially correlated between antennas, while small amplitude fluctuations occur over short periods and are uncorrelated between antennas, depending on the baseline length. This effect can be seen in Figure 4 by the fact that the antennas at the outer stations (W14 and W16) show larger amplitude fluctuations relative to the inner stations (W4 and W6). This occurs because the antenna-based phase solutions for each subarray are referenced to antennas at the center of the array, such that the reference antennas are within about 250 m of W4 and W6, but are separated from W16 and W14 by about 2000 m. An example of what occurs in the image plane due to tropospheric phase fluctuations is shown in Figure 5. Observations were made of the celestial calibrator 2007+404 at 22 GHz with the VLA at a resolution of 0.1$''$ (maximum baseline = 30 km) for a period of 1 hour. The data were self-calibrated using a long solution averaging time of 30 minutes, ie. just a mean phase was removed from each half of the data. `Snap-shot' images were then made from one minute of data at the beginning and end of the observation (upper left and upper right frames in Figure 5, respectively). Two important trends are apparent in these two frames. First, notice the positive-negative side-lobe pairs straddling the peak, indicative of antenna-based phase errors (Ekers 1998). These image artifacts are due to phase fluctuations which arise in small scale water vapor structures in the troposphere that are not correlated between antennas. Second, notice that the peak in each image has shifted from the true source position. This position shift is due to phase fluctuations which arise in large scale water vapor structures in the troposphere that are correlated between antennas, ie. a phase gradient across the array. These two frames are analogous to optical `speckle' images, although the timescales in the radio are much longer than in the optical due to the larger spatial scales for the turbulence. The lower left frame shows the image of 2007+404 made using the full hour of data, but with a self-calibration averaging time of 30 minutes. The source appears extended in this image. The lower right frame shows the same image after self-calibration with an averaging time of 30 seconds, and in this case the source is unresolved, ie. the source size is equivalent, within the noise, to the interferometric synthesized beam. The lower left image has a peak surface brightness of 1.0 Jy beam$^{-1}$, an off-source rms noise level of 47 mJy beam$^{-1}$, and a total flux density of 1.5 Jy. The lower right image has a peak surface brightness of 1.6 Jy beam$^{-1}$, an off-source rms noise level of 5 mJy beam$^{-1}$, and a total flux density of 1.6 Jy. Not correcting for tropospheric phase noise in the lower left frame has: (i) increased the off-source noise in the image, (ii) decreased the `coherence' (ie. lowered the peak surface brightness), and (iii) degraded the resolution (`seeing'). The important lesson from Figure 5 is that, while tropospheric phase errors can be quantified in terms of an antenna-based phase error, the errors are partially correlated between antennas on certain spatial and temporal scales, leading to positional shifts of sources as well as the standard positive-negative side-lobe pairs. Also, it is important to keep in mind that short baselines only `sample' the power in the phase screen on scales of order the baseline length. \section{Root Phase Structure Function} Tropospheric phase fluctuations are usually characterized by the spatial phase structure function, $\rm D_{\Phi}(b)$, \begin{equation} \rm D_{\Phi}(b) \equiv \langle ( \Phi(x + b) - \Phi(x) )^2 \rangle, \end{equation} where b is the distance between two antennas, $\rm \Phi(x + b)$ is the atmospheric phase measured at one antenna and $\rm \Phi(x)$ is the atmospheric phase measured at the second antenna, and the brackets represent an ensemble average. Usually in radio astronomy the ensemble average is replaced by a time average on one particular baseline. An interferometric array will sample the phase structure function at several baselines. For a single interferometer, the Taylor hypothesis (which asserts that temporal phase fluctuations are equivalent to spatial phase fluctuations) permits us to measure temporal phase fluctuations on a single baseline and translate these into the equivalent spatial phase structure function. In the following discussion we consider the square root of the phase structure function (the `root phase structure function'), which corresponds to the rms phase variations as a function of baseline length: $$\rm \Phi_{rms} \equiv \sqrt{D_{\Phi}}$$. Kolmogorov turbulence theory (Coulman 1990) predicts a function of the form: \begin{equation} \rm \Phi_{\rm rms}(b) = {K\over{\lambda_{mm}}} ~b^{\alpha} ~~ deg, \end{equation} where b is in km, and $\lambda$ is in mm. A typical value of K = 100 for the MMA site in Chajnantor under good weather conditions, and K = 300 is typical for the VLA site (Carilli, Holdaway, and Sowinski 1996, Sramek 1990). Kolmogorov turbulence theory predicts $\alpha$ = ${1\over3}$ for baselines longer than the width of the turbulent layer, W, and $\alpha$ = ${5\over6}$ for baselines shorter than W (Coulman 1990). The change in power-law index at b $=$ W is due to the finite vertical extent of the turbulent layer. For baselines shorter than W the full 3-dimensionality of the turbulence is involved (thick-screen), while for longer baselines a 2-dimensional approximation applies (thin-screen). Turbulence theory also predicts an `outer-scale', L$_{\circ}$, beyond which the rms phase variations should not increase with baseline length (ie. $\alpha$ = 0). This scale corresponds to the largest coherent structures, or maximum correlation length, for water vapor fluctuations in the troposphere, presumably set by external boundary conditions. Recent observations with the VLA support Kolmogorov theory for tropospheric phase fluctuations. Figure 6 shows the root phase structure function made using the BnA configuration of the VLA. This configuration has good baseline coverage ranging from 200m to 20 km, hence sampling all three hypothesized ranges in the structure function. Observations were made at 22 GHz of the VLA calibration source 0748+240. The total observing time was 90 min, corresponding to a tropospheric travel distance of 54 km, using v$_a$ = 10 m s$^{-1}$ (see section 6.2). The open circles show the nominal tropospheric root phase structure function over the full 90 min time range.\footnote{Note that the total observing time for calculating the rms phase fluctuations must be long for the larger configurations of the VLA, since the phase variations on a given baseline may have a significant, and perhaps even dominant, contribution from structures in the troposphere as large as five times the baseline length (Lay 1997).} The solid squares are the rms phases after subtracting (in quadrature) a constant electronic noise term of 10$^{\circ}$, as derived from the data by requiring the best power-law on short baselines. The 10$^{\circ}$ noise term is consistent with previous measurements at the VLA indicating electronic phase noise increasing with frequency as 0.5$^{\circ}$ per GHz (Carilli and Holdaway 1996). The three regimes of the structure function as predicted by Kolmogorov theory are verified in Figure 6. On short baselines (b $\le$ 1.2 km) the measured power-law index is 0.85$\pm$0.03 and the predicted value is 0.83. On intermediate baselines (1.2 $\le$ b $\le$ 6 km) the measured index is 0.41$\pm$ 0.03 and the predicted value is 0.33. On long baselines (b $\ge$ 6 km) the measured index is 0.1$\pm$0.2 and the predicted value is zero. The implication is that the vertical extent of the turbulent layer is:~ W $\approx$ 1 km, and that the outer scale of the turbulence is:~ L$_{\circ}$ $\approx$ 6 km. The increase in the scatter of the rms phases for baselines longer than 6 km may be due to an anisotropic outer scale (Carilli and Holdaway 1997). In practice, single baseline site testing interferometers tend to see values of $\alpha$ which form a continuous distribution between the theoretical thin and thick layer values of ${1\over3}$ and ${5\over6}$ (Holdaway et al., 1995). The distribution cuts off fairly sharply at these two theoretical values. Such a distribution could be due to the presence of both a thin turbulent layer associated with the ground or an inversion layer and a thick turbulent layer. By changing the relative weight of each of these two layers with their theoretical power law exponents, the resulting phase structure function will be very close to a power law with an exponent between the two theoretical values. Alternatively, the intermediate power law values could just reflect the transition between thin and thick turbulence. \section{Effects of Tropospheric Phase Noise} \subsection{Coherence} Tropospheric phase noise leads to a number of adverse effects on interferometric observations at mm wavelengths. First is the loss of coherence of a measured visibility on a given baseline over a given averaging time due to phase variations. For a given visibility, V = V$_{\circ}$e$^{i \phi}$, the effect on the measured amplitude due to phase noise in a given averaging in time is: \begin{equation} <\rm V> = \rm V_{\circ} \times <e^{i \phi}> = V_{\circ} \times e^{-\phi_{\rm rms}^2/2} \end{equation} assuming Gaussian random phase fluctuations with an rms variation of $\phi_{\rm rms}$ over the averaging time (Thompson, Moran, and Swenson 1986). For example, for $\phi_{\rm rms}$ = 1 rad, the coherence is:~ ${<\rm V>}\over{\rm V_{\circ}}$ = 0.60, meaning the observed visibility amplitude is reduced by 40$\%$ from the true value. \subsection{Seeing} A second effect of tropospheric phase fluctuations is to limit the spatial resolution of an observation in a manner analogous to optical seeing, where optical seeing is due to thermal fluctuations rather than water vapor fluctuations. Since interferometric phase corresponds to the measurement of the position of a point source (Perley 1998), it is clear that phase variations due to the troposphere will lead to positional variations of a source, and hence `smear-out' a point source image over time (Figure 5). The magnitude of tropospheric seeing can be calculated by considering the coherence as a function of baseline length. Since the coherence decreases for longer baselines given an averaging time long compared to the array crossing time, the observed visibility amplitude decreases with increasing baseline length, as would occur if the source were resolved by the array. Using equation 3 for the root phase structure, and equation 4 for the coherence, the visibility amplitude as a function of baseline length becomes: \begin{equation} <\rm V > = V_{\circ} \times exp(-[{{K' b^\alpha}\over{\lambda \sqrt{2}}}]^2) \end{equation} Note that the exponent must be in radians, so K$'$ = K $\times$ ${2 \pi}\over{360}$. The baseline length corresponding to the half-power point of the visibility curve, b$_{1/2}$, then becomes: $$ \rm b_{1/2} = (1.2 \times {{{\lambda_{mm}}\over{K'}}})^{1/\alpha} ~~km $$ For example, at 230 GHz using the typical value of $\alpha$ = 5/6, and a typical value for K$'$ at the MMA site of 1.7, the value of b$_{1/2}$ = 0.9 km. This means that the resolution of the array is limited by tropospheric seeing to: $\theta_{seeing}$ $\approx$ ${\lambda}\over{b_{1/2}}$ $\approx$ 0.3$''$ at 230 GHz. For average weather conditions, tropospheric seeing precludes diffraction limited resolution imaging for arrays larger than about 1 km at the MMA site in Chile, if no corrections are made for tropospheric phase noise. A rigorous treatment of tropospheric seeing, with predicted source sizes under various assumptions about the turbulence, can be found in Thompson, Moran, and Swenson (1986). Two important points need to be remembered when considering tropospheric seeing. First is that the root phase structure function flattens dramatically on baselines longer than $\approx$ 1 km, such that the tropospheric seeing degrades slowly with longer baselines. And second, there is an explicit connection between the coherence loss and the seeing: on the short baselines, where the phase errors are smaller, there is less coherence loss, and the correct flux density is measured even for a long averaging time. On the longer baselines, the phase errors are larger, causing decorrelation of the visibilities, either within an integration or across many integrations, thereby fictitiously resolving the source. It is the selective loss of coherence on the long baselines which determines the seeing. This phenomenon can be seen in the lower left frame of Figure 5, in which the peak surface brightness is only 60$\%$ of the expected peak, but the total flux density averaged over the `seeing disk' is 94$\%$ of the true value, ie. the shortest baselines see the total flux density of the source even for long averaging times. \subsection{Anomalous Refraction} A final problem arising from tropospheric phase variations is `anomalous refraction', or tropospheric induced pointing errors (Holdaway 1997, Butler 1997, Holdaway and Woody 1998). This effect corresponds to tropospheric seeing on the scale of the antenna itself. Phase gradients across the antenna change the apparent position of the source on a time scale $\approx$ ${D}\over{\rm v_a}$ $\approx$ 1 second, for an antenna with a diameter $D = 10$ m. A straight-forward application of Snells' law shows that the effect in arc-seconds should decrease with antenna diameter as $D^{\alpha - 1}$, or about $D^{-0.4}$ for median $\alpha$ of 0.6. The decrease in the pointing error with dish diameter is due to the fact that the value of $\alpha$ in the root phase structure function is less than unity, and hence the angle of the `wedge' of water vapor across the antenna becomes shallower with increasing antenna size. However, in terms of fractional beam size, the effect becomes worse with antenna size as $D^{\alpha}$. For the 10m MMA antennas the expected magnitude of the effect at an elevation of 50$^o$ is $\approx$ 0.6$''$, which is about 50$\%$ of the pointing error budget for the antennas. \section{Stopping the Troposphere: Techniques to Reduce the Effects of Tropospheric Phase Noise} An important point to keep in mind is that while tropospheric phase variations can be quantified in terms of a baseline-length dependent structure function, the errors are fundamentally antenna-based, and hence can be corrected by antenna-based calibration schemes, such as self-calibration or fast switching calibration. \subsection{Self-Calibration} A straight forward method of reducing phase errors due to the troposphere is self-calibration (Cornwell 1998). Self-calibration removes the baseline-dependent term in the root structure function, $\Phi_{\rm rms}$(b), leaving the residual tropospheric phase noise dictated by the `effective baseline': b$_{\rm eff}$ = ${\rm v_a t_{\rm ave}}\over2$ = half the distance the troposphere moves during the self-calibration averaging time, t$_{\rm ave}$. The factor of two arises from the fact that the mean calibration applies to the middle of the solution interval. The Taylor hypothesis dictates a relationship between temporal and spatial fluctuations such that the longer baselines will not sample the full power in the root phase structure function if the calibration cycle time is shorter than the baseline crossing time for the troposphere. Of course, we would like to make $\rm t_{\rm ave}$ as short as possible, but for a target source of some given brightness, we are limited in that we must detect the source in t$_{\rm ave}$ on each baseline with sufficient signal-to-noise ratio (SNR $\approx$ 2 for arrays with large numbers of antennas; Cornwell 1998) to be able to solve for the phase. Hence, there will be sources which are so weak that they cannot be detected in a time short enough to track the atmospheric phase fluctuations. For the MMA at 230 GHz, self-calibration should be possible on fairly weak continuum sources (of order 10 mJy), with fairly short integration times ($\approx$ 30 sec), leading to residual rms phase errors $\le$ 20$^o$. For the completed VLA 43 GHz system the limit is 50 mJy sources with 30 second averaging times with residual rms phase variations of 10$^o$. Self-calibration is not possible for weaker continuum sources, or for weak spectral line sources, or in the case where absolute positions are required. In these cases other methods must be employed to `stop' tropospheric phase variations. \subsection{Fast Switching} Another method for reducing tropospheric phase variations is `Fast Switching' (FS) phase calibration. This method is simply normal phase calibration using celestial calibration sources close to the target source (Fomalont and Perley 1998), only with a calibration cycle time, t$_{cyc}$, short enough to reduce tropospheric phase variations to an acceptable level (Holdaway 1992, Holdaway and Owen 1995, Carilli and Holdaway 1996, 1997). The expected residual phase fluctuations after FS calibration can be derived from the root phase structure function (equation 3), assuming an `effective baseline length', b$_{\rm eff}$, given by: \begin{equation} \rm b_{\rm eff}~ \approx~ d~ +~ {{v_a t_{cyc}}\over2} \end{equation} where $v_a$ = wind speed, and d = the physical distance in the troposphere between the calibrator and source. The FS technique will be effective for calibration cycle times shorter than the baseline crossing time of the troposphere = ${\rm b}\over{\rm v_{a}}$. Moreover, a significant gain is made when b$_{\rm eff}$ $<$ 1 km, thereby allowing for corrections to be made on the steep part of the root phase structure function (Figure 6), implying a timescale of 200 seconds or less for effective FS corrections. As with self-calibration, the calibrator source must be detected with sufficient SNR on time scales short enough to track the atmospheric phase fluctuations. The effectiveness of FS phase calibration is shown in Figure 7 for 22 GHz data from the VLA on baselines ranging from 100 m to 20 km. The solid squares show the nominal tropospheric root phase structure function averaged over 90 minutes (Figure 6). The open circles are the rms phases of the visibilities after applying antenna based phase solutions averaged over 300 seconds. The stars are the rms phases of the visibilities after applying antenna based phase solutions averaged over 20 seconds. The residual root structure function using a 300 second calibration cycle parallels the nominal tropospheric root structure function out to a baseline length of 1500m, beyond which the root structure function saturates at a constant rms phase value of 20$^{\circ}$. The implied wind velocity is then:~ v$_a$ = ${2 \times 1500 m}\over{300 sec}$ = 10 m s$^{-1}$. Using a 20 second calibration cycle reduces b$_{\rm eff}$ to only 100 m, which is shorter than the shortest baseline of the array, and the saturation rms is 5$^{\circ}$. The important point is that, after applying standard phase calibration techniques on timescales short compared the array crossing time of the troposphere, the resulting rms phase fluctuations are {\sl independent of baseline length for b $>$ b$_{\rm eff}$.} The FS technique allows for diffraction limited imaging of faint sources on arbitrarily long baselines. Note that this conclusion should also apply to Very Long Baseline Interferometry (VLBI), although the problems are accentuated due to the fact that antennas can be observing at very different elevations at any given time, and that VLBI observations typically employ low elevation observations in order to maximize mutual visibility times for widely separated antennas (see the discussion of the VLBI phase referencing technique by Beasley and Conway 1995). An important question to address when considering FS phase calibration is: are there enough calibrators in the sky in order to take advantage of a switching time as short as 40 seconds? This depends on the slew rate and settling time of the telescope, the set-up time of the electronics, the sensitivity of the array, and the sky surface density of celestial calibrators. Holdaway, Owen, and Rupen (1994) estimated the calibrator source counts at 90~GHz. They measured the 90~GHz flux densities of 367 flat spectrum quasars known from centimeter wavelength surveys, thereby determining the distribution of the spectral index between 5~GHz and 90~GHz, which was statistically independent of source flux density. By applying this spectral index distribution to well understood 5~GHz flat spectrum source counts, they were able to estimate the integral source counts for appropriate phase calibrators at 90~GHz. They estimate that the integral source counts over the whole sky between 0.1 to 1.0 Jy is about 170 $S^{3/2}_{90}$, where S$_{90}$ is the 90 GHz source flux density in Jy. Then, the typical distance to the nearest appropriate calibrator at 90~GHz will be about $\rm \theta \approx 7 \times S_{\nu}^{0.75}~~ deg$. For calibrators with S$_{\nu}$ $\ge$ 50 mJy, and assuming a slew rate of 2 deg s$^{-1}$, the 40 element MMA should be able to employ FS phase calibration on most sources with total cycle times $\le$ 20 sec, leading to residual rms phase fluctuations $\le$ 20$^o$ at 230 GHz, and on-source duty cycles $\approx$ 75$\%$. One important practical problem is the lack of all-sky surveys at high frequency from which to generate calibrator source lists. However, the MMA will be both agile and sensitive enough to survey a few square degrees around the target source in a few minutes prior to the observations, thereby permitting the determination of the optimal calibrator. \subsection{Paired Array Calibration} A third method for reducing tropospheric phase noise for faint sources is paired antenna, or paired array, calibration. Paired Array (PA) calibration involves phase calibration of a `target' array of antennas using a separate `calibration' array, where the target array is observing continuously a weak source of scientific interest while the calibration array is observing a nearby calibrator source (Holdaway 1992, Counselman et al., 1974, Asaki et al., 1996, 1998, Drashkik and Finkelstein 1979). In its simplest form PA calibration implies applying the phase solutions from a calibration array antenna to the nearest target array antenna at each integration time. An improvement can be made by interpolating the solutions from a number of nearby calibration array antennas to a given target array antenna at each integration time. Ultimately, the discrete measurements in space and time of the phases from the calibration array could be incorporated into a physical model for the troposphere to solve for, and remove, the effects of the tropospheric phase screen on the target source as a function of time and space using some intelligent method of data interpolation, such as forward projection using a physical model for the troposphere and Kalman filtering of the spatial time series (Zheng 1985). For simple pairs of antennas, the residual phase error can be derived from the root phase structure function with:~ $$\rm b_{\rm eff} \approx d + \Delta b$$ where d is the same as in equation 10, and $\Delta$b is the baseline length between the calibration antenna and the target source antenna. Figure 4 shows an observation for which PA calibration was implemented. Again, observations were made at 22 GHz using two `inter-laced' subarrays observing two close calibrators, 0432+416 and 0423+416. Notice how the phase variations for adjacent antennas in the different subarrays track each other closely. This correlation between phase variations from neighboring antennas in different subarrays observing different sources implies that the tropospheric phase variations can be corrected using PA calibration. In Figure 4 the temporal variations for neighboring antennas track each other well, but the mean phase over the observing time range is different between antennas. This phase off-set is due to the electronics and/or optics at each antenna, and should be slowly varying in time. Before interpolating phase solutions from the calibration array to the target array one must first determine, and remove, the electronic phase off-sets. This can be done by observing a celestial calibrator every 30 min or so. A demonstration of this process is shown in Figure 8. The upper frame in Figure 8 shows a random phase distribution along the west arm at a given time before the mean electronic phase is removed, while the lower frame in Figure 8 shows a smooth phase gradient along the arm after correction for the electronic phase term, indicating a tropospheric phase `wedge' down the array arm at this time. A quantitative measure of the effects of PA calibration can be seen in the root phase structure function plotted in Figure 9. The open triangles show the root phase structure function for the given observing day, as determined from the data with only the mean phase calibration (30 min averaging) applied. This function can be fit by a power-law in rms phase versus baseline length with index 0.65. The rms magnitude of 35$^o$ on a baseline of 1000m is somewhat higher than the expected value of about 25$^o$ on a typical summer evening at the VLA (Carilli etal. 1996). The stars in Figure 9 correspond to the `noise floor' for the phase measurements, as determined by calculating the root structure function from self-calibrated data with an averaging time of 30 seconds. The solid squares in Figure 9 show the root structure function for the data with PA calibration applied. The PA process in this case entailed interpolating phase solutions from neighboring antennas observing the `calibration source' 0423+418, to the antennas observing the `target source' 0432+416. The residual rms phase values are about 10$^o$ on short baselines, and increase very slowly with baseline length. These data indicate a significant improvement in rms phase fluctuations after application of PA calibration for baselines longer than about 300 m. The increased noise floor for the PA calibrated data relative to self-calibration indicates residual short-timescale phase differences which do not replicate between the target and calibration arrays. This noise floor is a combination of `jitter' in the electronic phase contribution, and residual tropospheric phase noise as determined by b$_{\rm eff}$ above. Note that the residual noise floor increases slowly with baseline length. This is due to the logarithmically increasing separation between VLA antennas along the arm. \section{Radiometry} The brightness temperature of atmospheric emission, T$_{\rm B}^{\rm atm}$, can be measured using a radiometer, and is given by the radiometry equation (Dicke et al., 1946): \begin{equation} \rm T_{\rm B}^{\rm atm}~ =~ \rm T_{\rm atm} ~ \times ~ (1~ -~ e^{-\tau_{\rm tot}}) \end{equation} where T$_{\rm atm}$ is the physical temperature of the atmosphere, and $\tau_{\rm tot}$ is the optical depth, which depends on, among other things, the precipitable water vapor content (PWV) of the troposphere. By measuring fluctuations in atmospheric brightness temperature with a radiometer, one can infer the fluctuations in the column density of water vapor of the troposphere (Barrett and Chung 1962, Staguhn et al., 1998, Staelin 1966, Westwater and Guiraud 1980, Rosenkranz 1989, Bagri 1994, Sutton and Hueckstadt 1997, Lay 1998). The relationship between electrical pathlength and water vapor column (equation 1) can then be used to derive the variable contribution from water vapor to the interferometric phase (Elgered 1993). This technique has been used with varying degrees of success at connected-element mm interferometers (Welch 1994, Woody and Marvel 1998, Bremer et al., 1997), and in geodetic VLBI experiments (Elgered et al. 1991). We assume that the atmospheric opacity can be divided into three parts: \begin{equation} \rm \tau_{\rm tot} ~ = ~ A_{\nu}\times w_{\circ}~ +~ B_{\nu}~ + ~A_{\nu}\times w_{\rm rms} \end{equation} where: (i) A$_{\nu}$ is the optical depth per mm of PWV as a function of frequency, (ii) w$_{\circ}$ is the temporally stable (mean) value for PWV of the troposphere, (iii) B$_\nu$ is the total optical depth due to dry air as a function of frequency (also assumed to be temporally stable), and (iv) w$_{\rm rms}$ is the time variable component of the PWV of the troposphere. It is this time variable component which causes the tropospheric phase `noise' for an interferometer. In effect, we assume a constant mean optical depth: $\rm \tau_{\circ} \equiv A_{\nu}\times w_{\circ}~ +~ B_{\nu}$, with a fluctuating term due to changes in PWV:~ $\rm \tau_{\rm rms} \equiv A_{\nu}\times w_{\rm rms}$, and that $\tau_{\circ}$ $>>$ $\tau_{\rm rms}$. Inserting equation 8 into equation 7, and making the reasonable assumption that A$_\nu$$\times$w$_{\rm rms} << 1$, leads to: \begin{equation} \rm T_{\rm B} ~ =~ T_{\rm atm}\times [1 - e^{-\tau_{\circ}}] ~~ + ~~ T_{\rm atm}\times e^{-\tau_{\circ}}\times[A_{\nu}\times w_{\rm rms} ~ + ~ {{(A_{\nu}\times w_{\rm rms})^2}\over{2}} ~ + ~ ...] \end{equation} The first term on the right-hand side of equation 9 represents the mean, non-varying T$_{\rm B}$ of the troposphere. The second term represents the fluctuating component due to variations in PWV, which we define as: \begin{equation} \rm T_{\rm B}^{\rm rms} ~ \equiv ~ T_{\rm atm}\times e^{-\tau_{\circ}}\times[A_{\nu}\times w_{\rm rms} ~ + ~ {{(A_{\nu}\times w_{\rm rms})^2}\over{2}} ~ + ~ ...] \end{equation} At first inspection, it would appear that equation 10 applies to fluctuations in a turbulent layer at the top of the troposphere, since the fluctuating component is fully attenuated (ie. multiplied by $e^{-\tau_{\circ}}$). However, for a turbulent layer at lower altitudes there is the additional term of attenuation of the atmosphere above the turbulent layer by the turbulence. It can be shown that the terms exactly cancel for an isobaric, isothermal atmosphere, in which case equation 10 is {\sl independent} of the height of the turbulence. Absolute radiometric phase correction entails measuring variations in brightness temperature with a radiometer, inverting equation 10 to derive the variation in PWV, and then using equation 1 to derive the variation in electronic phase along a given line of sight. As benchmark numbers for the MMA we set the requirement that we need to measure changes in tropospheric induced phase above a given antenna to an accuracy of ${\lambda}\over{20}$ at 230 GHz at the zenith, or $\phi_{\rm rms}$ = 18$^o$. This requirement inserted into equation 10 then yields a required accuracy of:~ w$_{\rm rms}$ = 0.01 mm. This value of w$_{\rm rms}$ then sets the required sensitivity, T$_{\rm B}^{\rm rms}$, of the radiometers as a function of frequency through equation 10. For the VLA we set the ${\lambda}\over{20}$ requirement at 43 GHz, leading to:~ w$_{\rm rms}$ = 0.05 mm. In its purest form, the inversion of equation 10 requires:~ (i) a sensitive, absolutely calibrated radiometer, (ii) accurate models for the run of temperature and pressure as a function of height in the atmosphere, and (iii) an accurate value for the height of the PWV fluctuations. Figure 10 shows the required sensitivity of the radiometer, T$_{\rm B}^{\rm rms}$, given the benchmark numbers for w$_{\rm rms}$ for the VLA and the MMA and using equation 10. It is important to keep in mind that lower numbers on this plot imply that more sensitive radiometry is required in order to measure the benchmark value of w$_{\rm rms}$. The required T$_{\rm B}^{\rm rms}$ values generally increase with increasing frequency due to the increase in A$_{\nu}$, with a local maximum at the 22 GHz water line, and minima at the strong O$_2$ lines (59.2 GHz and 118.8 GHz). The strong water line at 183.3 GHz shows a `double peak' profile, with a local minimum in T$_{\rm B}^{\rm rms}$ at the frequency corresponding to the peak of the line. This behavior is due to the product:~ A$_{\nu}$ $\times$ e$^{-\tau_{\circ}}$ in equation 10. The value of A$_{\nu}$ peaks at the line frequency, but this is off-set by the high total optical depth at the line peak. This effect is dramatic for the VLA case, where the required T$_{\rm B}^{\rm rms}$ at the 183 GHz line peak is very low. \subsection{Absolute Radiometric Phase Corrections} In this section we consider making an absolute correction to the electronic phase at a given antenna using an accurate, absolutely calibrated measurement of T$_{\rm B}$, and accurate measurements of tropospheric parameters (temperature and pressure as a function of height, and the scale height of the PWV fluctuations). We consider requirements on the gain stability, sensitivity, and on atmospheric data, given the benchmark values of w$_{\rm rms}$ and using equation 10 to relate w$_{\rm rms}$ and T$_{\rm B}^{\rm rms}$. We consider the requirements at a number of frequencies for the MMA site, including: (i) the water lines at 22.2 GHz and 183.3 GHz, (ii) the half power of the water line at 185.5 GHz, and (iii) two continuum bands at 90 GHz and 230 GHz. For the VLA we only consider the 22.2 GHz line. The results are summarized in Table 1. Row 1 shows the optical depth per mm PWV, A$_\nu$, at the different frequencies for the model atmospheres discussed in section 4, while row 2 shows the total optical depth, $\tau_{\rm tot}$, for the models. Row 3 shows the required T$_{\rm B}^{\rm rms}$ values as derived from equation 10. It is important to keep in mind that these values are simply the expected change in T$_{\rm B}$ given a change in w of 0.01 mm for the MMA and 0.05 mm for the VLA, for a single radiometer looking at the zenith. All subsequent calculations depend on these basic T$_{\rm B}^{\rm rms}$ values. The values range from 19 mK at 90 GHz, to 920 mK at 185.5 GHz, at the MMA site, and 120 mK for the VLA site at 22 GHz. We first consider sensitivity and gain stability. Row 4 lists approximate numbers for expected receiver temperatures, T$_{\rm rec+spill}$, in the case of cooled systems (eg. using the astronomical receivers for radiometry). Row 5 lists the contribution to the system temperature from the atmosphere, T$_{\rm rec,atm}$, and row 6 lists the expected total system temperature, T$_{\rm tot}$ (sum of row 4 and 5). Row 7 lists the rms sensitivity of the radiometers, T$_{\rm rms}$, assuming 1000 MHz bandwidth, one polarization, and a 1 sec integration time. In all cases the expected sensitivities of the radiometers are well below the required T$_{\rm B}^{\rm rms}$ values, indicating that sensitivity should not be a limiting factor for these systems. Row 8 lists the required gain stability of the system, defined as the ratio of total system temperature to T$_{\rm B}^{\rm rms}$:~ $\delta$Gain $\equiv$ $\rm {T_{\rm tot}}\over{T_B^{\rm rms}}$. Values range from 210 for the 185.5 GHz measurement to 5800 for the 90 GHz measurement at the MMA, and 450 for the VLA site at 22 GHz. Rows 9 and 10 list total system temperatures and expected rms sensitivities in the case of uncooled radiometers. We adopt a constant total system temperature of T$_{\rm tot}$ = 2000 K, but the other parameters remain the same (bandwidth, etc...). The radiometer sensitivity is then 63 mK in 1 second. This sensitivity is adequate to reach the benchmark T$_{\rm B}^{\rm rms}$ values in row 3, although at 230 GHz the sensitivity value is within a factor two of the required T$_{\rm B}^{\rm rms}$. The required gain stabilities in this case are listed in row 11. The requirement becomes severe at 230 GHz ($\delta$Gain = 15000). We next consider the requirements on atmospheric data, beginning with T$_{\rm atm}$. The dependence of T$_{\rm B}^{\rm rms}$ on T$_{\rm atm}$ comes in explicitly in equation 10 through the first multiplier, and implicitly through the effect of T$_{\rm atm}$ on $\tau_{\circ}$. For simplicity, we consider only the explicit dependence, which will lead to an under-estimate of the expected errors by at most a factor $\approx$ 2 -- adequate for the purposes of this document (Sutton and Hueckstaedt 1997). Under this simplifying assumption the required accuracy, $\delta$T$_{\rm atm}$, becomes: $$\rm \delta T_{\rm atm} ~ \approx ~ {{T_B^{\rm rms}}\over{[1-e^{-\tau_{\circ}}]}} ~ K$$ The values $\delta$T$_{\rm atm}$ are listed in row 12. Values in parentheses are the percentage accuracy in terms of the ground atmospheric temperature. Values are typically of order 1 K, or a few tenths of a percent of the mean. A related requirement is the accuracy of the gradient in temperature:~ $\delta$$\rm {dT}\over{dh}$ $\approx$ $\rm {\delta T_{\rm atm}}\over{h_{\rm turb}}$, ~ in the case of a turbulent layer at h$_{\rm turb}$ = 2 km, and assuming a very accurate measurement of T$_{\rm atm}$ on the ground and a very accurate measurement of h$_{\rm turb}$. These values are listed in row 13. The accuracy requirements range from 0.25 K km$^{-1}$ to 1.5 K km$^{-1}$, or roughly 10$\%$ of the mean gradient. Similarly, we can consider the required accuracy of the measurement of the height of the troposphere, $\delta$h$_{\rm turb}$ $\approx$ $\rm {\delta T_{\rm atm}}\over{{{dT}\over{dh}}}$, assuming a perfect measurement of the ground temperature and temperature gradient. These values are listed in row 14. Values are typically a few tenths of a km, or roughly 10$\%$ of h$_{\rm turb}$. We consider the requirements on atmospheric pressure given the T$_{\rm B}^{\rm rms}$ requirements. The relationship between T$_{\rm B}$ and w is affected by atmospheric pressure through the change in the pressure broadened line shapes. An increase in pressure will transfer power from the line peak into the line wings, thereby flattening the overall profile. The expected changes in optical depth (or brightness temperature) as a function of frequency have been quantified by Sutton and Hueckstaedt (1997), and their coefficients relating changes in pressure with changes in optical depth are listed in row 17. Note the change in sign of the coefficient on the line peaks versus off-line frequencies. Sutton and Hueckstaedt point out that, since the integrated power in the line is conserved, there are `hinge points' in the line profiles where pressure changes have very little effect on T$_{\rm B}$, ie. for an increase in pressure at fixed total PWV the wings of the line get broader while peak gets lower. These hinge points are close to the half power points in T$_{\rm B}$ of the lines. Rows 15 and 16 list the requirements on the accuracy of P$_{\rm atm}$, and on the value of h$_{\rm turb}$. The values of $\delta$P$_{\rm atm}$ are derived from the equation:~ $\delta$P$_{\rm atm}$ = $\rm {A_{\nu}w_{\rm rms}}\over{\tau_{\rm tot}X}$, where $\rm X$ is the coefficient listed in row 17. We find that the value of P$_{\rm atm}$ needs to be known to about 1$\%$, and the height of the turbulent layer needs to be known to a few percent. The exception is at the hinge point of the line ($\approx$ 185.5 GHz), where the optical depth is nearly independent of P$_{\rm atm}$. There are a few potential difficulties with absolute radiometric phase corrections which we have not considered. First is the question of how to make a proper measurement of the `ground temperature'? It is possible, and perhaps likely, that the expected linear temperature gradient of the troposphere displays a significant perturbation close to the ground. The method for making the `correct' ground temperature measurement remains an important issue to address in the context of absolute radiometric phase correction. Second, we have only considered a simple model in which the PWV fluctuations occur in a narrow layer at some height h$_{\rm turb}$, which presumably remains constant over time. If the fluctuations are distributed over a large range of altitude then one needs to know the height of the dominant fluctuation at {\sl each time} to convert T$_{\rm B}$ into electrical pathlength. And when fluctuations at different altitudes contribute at the same time, this conversion becomes problematical. Again, the required accuracies for the height of the fluctuations are given in rows 14 and 16 in Table 1. A possible solution to this problem is to find a linear combination of channels for which the effective conversion factor is insensitive to altitude under a range of conditions, ie. hinge points generalized to a multi-channel approach (Lay 1998, Staguhn et al., 1998). And third, the shape of the pass band of the radiometer needs to be known very accurately in order to obtain absolute T$_{\rm B}$ measurements. A final uncertainty involved in making absolute radiometric phase corrections are errors in the theoretical atmospheric models relating w and T$_{\rm B}$ (Elgered 1993). Sutton and Hueckstaedt (1997) point out that model errors are by far the dominant uncertainties when considering absolute radiometric phase correction, and they have calculated a number of models with different line shapes and different empirically determined water vapor continuum `fudge-factors'. Row 18 in Table 1 lists the approximate differences between the various models at various frequencies. Models can differ by up to 3 mm in PWV, corresponding to 19.5 mm in electrical pathlength, or 30$\pi$ rad in electronic phase at 230 GHz. The differences are most pronounced in the continuum bands, but are only negligible close to the peak of the strong 183 GHz line. This is an area of very active research, and it may be that this uncertainty is greatly reduced in the near future (Rosenkranz 1998). Given the status of current atmospheric models, radiometric phase correction then requires some form of empirical calibration of the water vapor continuum contribution in order to relate T$_{\rm B}$ to w. The exception may be a measurement close to the peak of the 183 GHz line, but in this case saturation becomes a problem. Perhaps most importantly, if the calibrated continuum term is due to incorrect line shapes, it will depend on both T$_{\rm atm}$ and P$_{\rm atm}$, in which case the continuum term may require frequent calibration. Overall, absolute radiometric phase correction requires: (i) systems that are sensitive (19 $\le$ T$_{\rm rms}$ $\le$ 920 mK), and stable over long timescales (200 $\le$ $\delta$Gain $\le$ 15000), and (ii) knowledge of the tropospheric parameters, such as T$_{\rm atm}$, P$_{\rm atm}$, and h$_{\rm turb}$, to a few percent or less. And even if such accurate measurements are available, fundamental uncertainties in the atmospheric models relating T$_{\rm B}$ and PWV may require empirical calibration of the T$_{\rm B}^{\rm rms}$ - w$_{\rm rms}$ relationship at regular intervals. Lay (1998) has recently presented an interesting radiometric phase correction method using multifrequency measurements of the 183 GHz water line profile. His method is insensitive to atmospheric parameters, since it relies on using the line profile, and in particular, the hinge points of the lines. This method may allow for an absolute radiometric phase correction to be made without great uncertainties due to the atmospheric models. \subsection{Empirically Calibrated Radiometric Phase Corrections} Many of the uncertainties in Table 1 arise from the fact that we are demanding an absolute phase correction at each antenna based on the measured T$_{\rm B}$ plus ancillary data (T$_{\rm atm}$, P$_{\rm atm}$, h$_{\rm turb}$,...), using a theoretical model of the atmosphere to relate T$_{\rm B}$ to w. This sets very stringent demands on the absolute calibration, on the accuracy of the ancillary data, and on the accuracy of the theoretical model atmosphere. The current atmospheric models under-predict w by large factors in the continuum bands, thereby requiring calibration of (possibly time dependent) water vapor continuum `fudge-factors'. One way to avoid some of these problems is to calibrate the relationship between fluctuations in T$_{\rm B}^{\rm rms}$ with fluctuations in antenna-based phase, $\phi_{\rm rms}$, by observing a strong celestial calibrator at regular intervals. This empirically calibrated phase correction method would circumvent dependence on ancillary data and model errors (Woody and Marvel 1998), and mitigate long term gain stability problems in the electronics. This technique can be thought of as calibrating the `gain' of both the atmosphere and the electronics, in terms of relating T$_{\rm B}^{\rm rms}$ to $\phi_{\rm rms}$. In its simplest form, empirically calibrated radiometric phase correction would be used only to increase the coherence time on source. No attempt would be made to connect the phase of a celestial calibrator with that of the target source using radiometry, and hence the absolute phase on the target source would still be obtained from the calibration source. Such a process is being implemented at the Owen Valley Radio Observatory (Woody and Marvel 1998). In this case the absolute phase is obtained from the first accurate phase measurement on the celestial calibrator, while the subsequent time series of phase measurements on the calibrator are then used to derive the T$_{\rm B}^{\rm rms}$ to $\phi_{\rm rms}$ relationship. This process results in additional phase uncertainty in a manner analogous to Fast Switching phase calibration (Holdaway and Owen 1995). The residual error is set by the distance between the calibrator and source, and the time required to obtain the first accurate record: t$_{cal} \equiv$ (the slew time + the integration time required for the first accurate phase measurement). In this case: $$\rm b_{eff}~ \approx~ v_{a} \times t_{cal}~ + ~d,$$ where d is the physical distance in the troposphere set by the angular separation of the calibrator and the source, and b$_{eff}$ is the `effective baseline' to be inserted into equation 3 in order to estimate the residual uncertainty in the absolute phase. For example, assuming t$_{cal}$ = 10 sec, and the calibrator-source separation = 2$^o$, leads to b$_{eff}$ = 170 m, or $\phi_{\rm rms}$ = 20$^o$ at 230 GHz at the MMA site. Note that the temporal character of this `phase noise' is unusual in that the short timescale (t $<<$ t$_{cyc}$) variations are removed by radiometry, while the long timescale variations (t $>>$ t$_{cyc}$) are removed by celestial source calibration (Lay 1997). An example of such a relatively calibrated radiometric phase correction is shown in Figure 11, using data from the VLA at 22 GHz. Observations were made of the celestial calibrator 0319+415 (3C 84). In the upper frame, the dash line shows the interferometric phase time series measured between antennas 5 and 9, corresponding to a baseline length of about 3 km. The solid line shows the predicted phase time series derived by differencing measurements of the 22 GHz system temperature at each antenna. A single scale factor relating phase fluctuations and fluctuations in system temperature differences was derived from all the data, by requiring a minimum residual rms scatter in the phase fluctuations after applying radiometric phase correction. A constant off-set was also applied to each data set. Note the clear correlation between measured interferometric phase variations, and the phase variations predicted by radiometry. The middle frame shows the residual phase variations after radiometric correction. The rms variations in the raw phase time series before correction are 32$^o$. After applying the radiometric correction, the rms phase variations are reduced to 17$^o$. The lower frame shows the residual phase variations after radiometric correction, but now making a correction for the first and second half of the data separately. The residual rms phase variations are now 13$^o$. The scale factor changes by about 10$\%$ over the 36 minutes. This `empirically calibrated' radiometric phase correction technique has been implemented successfully at the Owens Valley Radio Observatory and at the IRAM interferometer (Woody and Marvel 1998, Bremer et al., 1997). A number of questions remain to be answered concerning this technique, including: (i) over what time scale and distance will this technique allow for radiometric phase corrections when switching between the source and the calibrator? And (ii) how often will calibration of the T$_{\rm B}^{\rm rms}$ - $\phi_{rms}$ relationship be required, ie. how stable are the radiometers and the mean parameters of the atmosphere? \subsection{Clouds and Other Issues} Water droplets present the problem that the drops contribute significantly to the measured T$_{\rm B}$ but not to w, thereby invalidating the model relating T$_{\rm B}$ and w. This problem can be avoided by using multichannel measurements around the water lines (183 GHz or 22 GHz), since T$_{\rm B}$ for the lines is not affected by water drops. Alternatively, a dual-band system could be used to separate the effect of water drops from water vapor (eg. 90 GHz and 230 GHz), since the frequency dependence of T$_{\rm B}$ is different for the two water phases. This later method requires a multi-band radiometer, which may be difficult within the context of the MMA antenna design. The question of whether clouds will be a significant problem on high quality sites such as the MMA site in Chile remains to be answered. A final error introduced when using radiometric phase calibration is due to the different path through the troposphere seen by the radiometer with that seen by the astronomical receiver. If the radiometer is not the astronomical receiver itself, then the angular separation of the radiometer beam and the telescope beam, $\theta_{diff}$, can be of order a few degrees, depending on the lay-out of the receivers and the telescope optics. This angle corresponds to 100 m or so at a height of 2 km. The magnitude of the error introduced by such an observing path difference can be derived from the root phase structure function assuming b$_{eff}$ = h$_{\rm turb}$ $\times$ $\theta_{diff}$. \section{Fast Switching vs. Paired Array Calibration vs. Radiometric Phase Correction} Fast Switching phase calibration has been used extensively, and successfully, at the VLA for diffraction limited imaging of faint astronomical sources on 30 km baselines at 43 GHz (Lim et al., 1998, Wilner et al., 1996). However, the required switching times at higher frequencies (10's of seconds or less) has thus far precluded the use of FS for existing mm observatories. Paired Array calibration has been demonstrated effective in reducing tropospheric phase noise for interferometers (Asaki et al., 1996, 1998), but has not yet been implemented as a standard technique for astronomical observing. Radiometric phase correction has been used to varying degrees of success at different observatories, but mostly on an experimental basis (Welch 1994, Bremer 1997, Woody and Marvel 1998, Elgered et al., 1991). We summarize the relative advantages and disadvantages of the three techniques. The advantages of FS are that: (i) FS uses the full array to observe the target source, and (ii) FS removes the long and short term electronic phase noise along with the tropospheric phase noise. The disadvantages are that: (i) FS places stringent constraints on telescope design in terms of slew rate, mechanical settling time, and electronic set-up time, and (ii) on-source observing time is lost due to frequent moves and calibration. Simulations for the MMA (Holdaway 1998) indicate that overall fast switching efficiencies (including both sensitivity losses due to decreased observing time and increased decorrelation) of about 0.75 should be possible, assuming a distribution of observing frequencies from 30~GHz to 650~GHz and using the measured distribution of atmospheric phase conditions at the Chajnantor site. Typical switching times range from 10 to 20 seconds. The advantages of PA calibration are that: (i) the `target array' observes the source continuously, and (ii) the demands on the antenna mechanics and electronics are less stringent than for FS. The disadvantages are that: (i) the electronic phase noise is not removed, (ii) the geometry of the array must allow for neighboring antennas, even in large (sparsely populated) arrays, and (iii) the number of visibilities from the target source array decreases quadratically with the decreasing number of antennas. These latter two effects can reduce significantly the Fourier spacing coverage and sensitivity of the array. One possible solution to these problems is to have a `calibration array' of smaller, cheaper antennas strategically placed with respect to the antennas of the main array which is dedicated to tropospheric phase calibration by observing celestial calibrators at a fixed frequency (eg. 90 GHz). The solutions would then be extrapolated to the observing frequency of the main array using the linear relationship between tropospheric phase and frequency (equation 1). This would require a separate correlator and IF system, and complications may arise due to the fact that the wet troposphere becomes dispersive at frequencies higher than about 400 GHz (Sutton and Hueckstaedt 1997). For PA and FS calibration, if the bulk of the phase fluctuations occur in a thin turbulent layer, it may be possible to perform an intelligent method of data interpolation, such as forward projection using a physical model for the troposphere and Kalman filtering of the spatial time series, to account for the motion of the atmosphere across the array. The advantages of radiometric phase correction are that it: (i) alleviates constraints on telescope mechanics and array design, and (ii) the full array can observe the target source continuously. The disadvantages are that: (i) it places constraints on receiver lay-out such that the radiometer is always looking at the sky, (ii) it does not remove the electronic phase noise, and (iii) there remains significant questions about the viability of absolute radiometric phase correction, or the limitations imposed when using an `empirically calibrated' radiometric phase correction technique (see sections 7.1 and 7.2). All three calibration methods are being investigated for the MMA, and we envision that all the methods will be employed to some degree at the MMA, depending on the configuration, the observing conditions, and the scientific requirements of a given observation. \footnotesize\acknowledgments We thank F. Owen, O. Lay, E. Sutton, and J. Carlstrom for useful comments on sections of this paper, and K. Desai for allowing us to use a figure from a previous publication. This research made use of the NASA/IPAC Extragalactic Data Base (NED) which is operated by the Jet propulsion Lab, Caltech, under contract with NASA. The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. \clearpage \newpage
\section{Introduction} According to the holographic principle \cite{hoof,suss1,suss2}, the bulk theory (with gravity) and the boundary theory (without gravity) are equivalent. Such correspondence gives insight into one theory from the other. Interest in the holographic principle is revived from the recent observation \cite{mald1} that the bulk gravity theory in the near horizon geometry of brane configuration is equivalent to the boundary theory described by the corresponding worldvolume field theory in the decoupling limit. (The closely related previous works in Refs. \cite{duff1,town1,town2} study the correspondence between the bulk theory on the AdS space and the boundary supersingleton field theory.) In particular, when the near horizon geometry is the AdS space, i.e. D3-brane, M2-brane and M5-brane cases, the boundary theory is conformal. However, it is found out in Refs. \cite{jevi1,jevi2} that even for other branes, whose near horizon geometry is not the AdS type and therefore the corresponding boundary theory is not a ``genuine'' conformal theory, one can still define ``generalized'' conformal theory. The symmetry group of such generalized conformal theory can be manifestly seen in the bulk theory in the so-called ``dual'' frame \cite{duff2,town1,town2,boon1,boon2,town3}. The dual frame is regarded as a preferred frame for supergravity probes in the near horizon background of the source branes. In this frame, the near horizon geometry of a $p$-brane supergravity solution takes the AdS$_{p+2}\times S^n$ form. The $SO(p+1,2)$ isometry of the AdS$_{p+2}$ part is realized in the boundary theory as the ``generalized'' conformal symmetry of Refs. \cite{jevi1,jevi2}. The dual frame is also called a ``holographic'' frame, since an UV/IR connection between the bulk and the boundary theory is manifest in this frame. In this paper, we study the D0-brane case of such generalized conformal theory. D0-brane is particularly interesting for its relevance to M theory. An important discovery of the M(atrix) model \footnote{This model was proposed \cite{halp1} long ago as the $N=16$ supersymmetric gauge quantum mechanics. See also Ref. \cite{halp2}.} is that difficult problems of quantum M theory is reduced to non-relativistic quantum mechanics. Originally, it is conjectured \cite{matr1} that M-theory in the infinite momentum frame (IMF) \cite{wein} is exactly described by the $U(\infty)$ $D=1$ super-Yang-Mills (SYM) theory, which is the worldvolume theory of the bound state of the infinite number ($N\to\infty$) of D0-branes. The $D=1$ $U(N)$ SYM theory, which is the supersymmetric matrix model description of the supermembrane theory \cite{dewi}, is nothing but the maximally supersymmetric $U(N)$ Yang-Mills theory dimensionally reduced from $9+1$ to $0+1$ dimensions (i.e. $N\times N$ hermitian matrix quantum mechanics). It is further conjectured \cite{matr2} that the equivalence of (M)atrix theory to M-theory is also valid for a finite $N$. The conjecture states that M-theory compactified on a light-like circle with finite momentum along the circle is exactly described by a $U(N)$ matrix theory. The quantization of such theory is called the discrete light-cone quantization (DLCQ). One can think of this light-like circle as a small space-like circle that is boosted by a large amount \cite{seib1}. Just like M-theory in the IMF, M-theory in the light-cone frame (LCF) is described purely by D0-branes with positive momentum and has the Galilean invariance in the transverse space \cite{matr2}. Therefore, M-theory on a light-like circle can be viewed as a theory of the finite number of D0-branes in the low velocity (non-relativistic) limit. This idea can also be understood from the bulk/boundary duality as follows. When compactified on a light-like circle, the supergravity M-wave solution (viewed as the supergraviton with the momentum number $N$) becomes the near horizon limit of supergravity solution for $N$ coinciding D0-branes \cite{tsey1}. Since M-theory in the LCF in the supergravity solution level is related to the near horizon limit of the supergravity D0-brane solution, by the generalized AdS/CFT duality \cite{mald2,jevi1,jevi2} M-theory in the LCF has to be related to the boundary theory of the bound state of $N$ D0-branes. This boundary theory is the maximally supersymmetric $SU(N)$ Yang-Mills theory dimensionally reduced from $9+1$ to $0+1$ dimensions, as stated in the (M)atrix theory conjecture. In this paper, we view the non-relativistic (matrix) quantum mechanics of bound states of D0-brane as a system of the probe D0-brane moving in the background of large number of source D0-brane bound state. This view of (M)atrix model is also taken in Ref. \cite{tsey1}, which reproduces the (M)atrix model graviton-graviton scattering calculation from the effective action for a probe moving in the background of the M-wave supergravity solution. It is further shown in Refs. \cite{jevi1,jevi2} that the Dirac-Born-Infeld (DBI) action for a radially moving probe D0-brane in the background of the source D0-brane bound state can also be determined by imposing the generalized conformal symmetry of the boundary theory, which is just the $D=1$ $U(N)$ SYM theory describing the M theory in the LCF. The quantum mechanics of the probe D0-brane in the near horizon background of the source D0-brane is a reminiscence of the conformal quantum mechanics of the charged test particle moving in the near horizon background of the $D=4$ extreme Reissner-Nordstr\"om black hole studied in Ref. \cite{kall1}. The radial motion of such test particle is described by the relativistic conformal quantum mechanics with the $SL(2,{\bf R})\cong SU(1,1)$ symmetry. This is a generalized version of (non-relativistic) conformal mechanics studied in Refs. \cite{scm1,ap,scm2}. One can view such generalized conformal mechanics of the radial motion of the test particle as the boundary conformal field theory counterpart of the bulk gravity theory in the near horizon Ads$_2$ geometry of the extreme Reissner-Nordstr\"om black hole, since the $SO(1,2) \cong SU(1,1)$ isometry of the AdS$_2$ space is realized as the symmetry of the dynamics of the test particle in the spacetime with one lower dimension. Since the near horizon geometry of the source D0-brane supergravity solution in the dual frame is AdS$_2\times S^8$, one would expect that the dynamics of the probe D0-brane in this background has the $SL(2,{\bf R})\cong SO(2,1)$ symmetry. In the D0-brane case, since this symmetry does not extend to the genuine conformal symmetry of the boundary theory but only to the so-called generalized conformal symmetry \cite{jevi1,jevi2} as pointed out in the above, the quantum mechanics of the probe D0-brane will have a generalized conformal symmetry. The paper is organized as follows. In section 2, we summarize the properties of the near horizon geometry of the supergravity D0-brane solution. In section 3, we study the ``generalized'' conformal quantum mechanics of D0-brane, elaborating its relation to M theory as pointed out in the above. \section{$D0$-brane Solution in the Near Horizon Region} In this section, we survey aspects of the $D0$-brane supergravity solution and its near horizon geometry, illuminating their relations. The string-frame type-IIA effective supergravity action for $D0$-brane solution is given by \begin{equation} S={1\over{16\pi G_{10}}}\int d^{10}x\sqrt{-G^{str}} [e^{-2\phi}({\cal R}^{str}+2\partial_M\phi\partial^M\phi) -{1\over{2\cdot 2!}}F_{MN}F^{MN}], \label{iibsg} \end{equation} where $G_{10}$ is the 10-dimensional gravitational constant, ${\cal R}^{str}$ and $G^{str}$ are the Ricci scalar and the determinant for the the string-frame metric tensor $G^{str}_{MN}$, $\phi$ is the dilaton in the NS-NS sector and $F_{MN}$ is the field strength for the RR 1-form potential $A_M$ that $D0$-branes couple to. The supergravity solution for the $D0$-brane has the following form: \begin{eqnarray} ds^2_{str}&=&G^{str}_{MN}dx^Mdx^N=-H^{-{1\over 2}}dt^2+H^{1\over 2} (dx^2_1+\dots+dx^2_9), \cr e^{\phi}&=&g_{s}H^{3\over 4},\ \ \ A_t=-g^{-1}_{s}H^{-1};\ \ \ H=1+{Q\over{r^7}}\equiv 1+\left({{\mu}\over r}\right)^7, \label{sgdzero} \end{eqnarray} where $g_s$ is the string coupling constant, which is just the vacuum expectation value or the asymptotic value of the dilaton $e^{\phi}$. Here, the $D0$-brane charge $Q$ is related to the string theory quantities as \begin{equation} Q=(\alpha^{\prime})^{7\over 2}g_sN, \label{charge} \end{equation} where $\alpha^{\prime}$ is related to the string length scale $l_s$ as $l_s=\sqrt{\alpha^{\prime}}$ and $N$ is the number of the $D0$-branes. By expanding the Dirac-Born-Infeld (DBI) action for the D0-brane to the lowest order in $\alpha^{\prime}$, one finds the following relation of the string theory quantities to the Yang-Mills gauge coupling $g_{YM}$: \begin{equation} g^2_{YM}={{g_s}\over{(\alpha^{\prime})^{3\over 2}}}. \label{ymcouple} \end{equation} Therefore, the constant $Q$ in the harmonic function is rewritten in terms of the SYM quantities as $Q=\mu^{7}=(\alpha^{\prime})^5g^2_{YM}N$. In order to decouple the massive string modes (whose masses are proportional to $1/\alpha^{\prime}$) and the gravity modes (whose strength goes as $g^2_s(\alpha^{\prime})^4$) from the massless open string modes, which describe the Yang-Mills theory, one has to take $\alpha^{\prime}\to 0$, while keeping $g_{YM}$ as a finite constant, which means for the D0-brane case $g_s\to 0$ as well (Cf. Eq. (\ref{ymcouple})). Furthermore, the above metric (\ref{sgdzero}) can be regarded as the gravitational field felt by a probe D0-brane in the background of the collection of $N$ numbers of source D0-branes with the radial coordinate $r$ being interpreted as the distance between and probe and source $D0$-branes \cite{mald3}. So, in order to keep the mass $m_{string}=r/\alpha^{\prime}$ of the state of open string, which stretches between the source and the probe, finite, one has also take the limit $r\to 0$ while keeping the following combination as a finite constant: \begin{equation} U\equiv {r\over{\alpha^{\prime}}}, \label{higgs} \end{equation} thereby going to the near horizon region of the supergravity solution (\ref{sgdzero}). This combination also corresponds to the conventional super-Yang-Mills scalar $\Phi^I=X^I/l^2_s$, whose vacuum expectation value sets the energy scale. In this decoupling limit, the supergravity solution (\ref{sgdzero}) takes the following form: \begin{eqnarray} ds^2_{str}&=&-\left({r\over\mu}\right)^{7\over 2}dt^2+ \left({\mu\over r}\right)^{7\over 2}(dr^2+r^2d\Omega^2_8) \cr &=&\alpha^{\prime}\left[-{{U^{7\over 2}}\over{g_{YM}\sqrt{N}}}dt^2 +{{g_{YM}\sqrt{N}}\over{U^{7\over 2}}}(dU^2+U^2d\Omega^2_8)\right], \cr e^{\phi}&=&g_{s}\left({\mu\over r}\right)^{{21}\over 4}=g^2_{YM} \left({{g^2_{YM}N}\over{U^7}}\right)^{3\over 4}, \cr A_t&=&g^{-1}_{s}\left({r\over\mu}\right)^7={\sqrt{\alpha^{\prime}} \over{g^2_{YM}}}{{U^7}\over{g^2_{YM}N}}. \label{sgdecoupl} \end{eqnarray} The above supergravity solution (\ref{sgdecoupl}) can be trusted when the spacetime curvature ${\cal R}\sim U^3/(g^2_{YM}N)$ is much smaller than the string scale $(\alpha^{\prime})^{-1}$ and string coupling $e^{\phi}$ is very small, leading to the following constraints on $U$ and the original radial coordinate $r$: \begin{equation} g^{2\over 3}_{YM}N^{1\over 7}\ll U \ll g^{2\over 3}_{YM}N^{1\over 3}, \label{ucnstrnt} \end{equation} \begin{equation} \sqrt{\alpha^{\prime}}g^{1\over 3}_{s}N^{1\over 7}\ll r \ll \sqrt{\alpha^{\prime}}g^{1\over 3}_{s}N^{1\over 3}. \label{rcnstrnt} \end{equation} On the other hand, the near horizon condition $r\ll\mu=Q^{1\over 7}$ is expressed in terms of string theory quantities as $r\ll \sqrt{\alpha^{\prime}}(g_sN)^{1\over 7}$. So, for sufficiently large $N$ and small $g_s$, the supergravity solution (\ref{sgdecoupl}) can be trusted in the overlapping region of $U$. Note, the near horizon geometry of the D0-brane supergravity solution in the string frame is not AdS$_2\times S^8$, since when the metric is expressed in the following suggestive form \begin{equation} ds^2_{str}=\alpha^{\prime}\left[-{{U^2}\over\sqrt{\rho_0}}dt^2+ \sqrt{\rho_0}{{dU^2}\over{U^2}}+\sqrt{\rho_0}d\Omega^2_8\right], \label{adsform} \end{equation} the radius $\rho_0\equiv Q/((\alpha^{\prime})^5U^3)$ of the would-be AdS space depends on the coordinate $U$. However, with the suitable choice of frame, called the ``dual'' frame \cite{duff2,town1,town2,town3}, the spacetime metric in the near-horizon region takes the AdS$_2\times S^8$ form. Namely, if one applies the Weyl transformation \footnote{In this paper, we do not include a factor involving $N$ in the Weyl transformation of the metric, unlike Ref. \cite{town3}, so that the metric in the standard AdS form in the horospherical coordinates has explicit dependence on $N$.} on the metric as $G^{str}_{MN}\to G^{dual}_{MN}=e^{-{2\over 7}\phi} G^{str}_{MN}$, then the metric in (\ref{sgdecoupl}) transforms to \begin{eqnarray} ds^2_{dual}&=&G^{dual}_{MN}dx^Mdx^N=g^{-{2\over 7}}_{s}\left[ -\left({r\over\mu}\right)^5dt^2+\left({{\mu}\over r}\right)^2dr^2+ \mu^2d\Omega^2_8\right] \cr &=&N^{2\over 7}\alpha^{\prime}\left[-{{U^{5}}\over{g^{2}_{YM}N}}dt^{2} +{{dU^{2}}\over{U^{2}}}+d\Omega^{2}_{8}\right]. \label{dualmet} \end{eqnarray} By further redefining the radial coordinate as $\bar{r}=({{25}\over 4} g^{4\over 7}_{s}\mu^3)^{-{1\over 2}}r^{5\over 2}$ or $u=({{25}\over 4} g^{2}_{YM}N^{3\over 7})^{-{1\over 2}}U^{5\over 2}$, one can bring the metric to the following standard AdS$_2\times S^8$ form in the horospherical coordinates: \begin{eqnarray} ds^2_{dual}&=&-\left({{5g^{1\over 7}_{s}}\over{2\mu}}\right)^2\bar{r}^2dt^2+ \left({{2\mu}\over{5g^{1\over 7}_{s}}}\right)^2{{d\bar{r}^2}\over{\bar{r}^2}} +\left({\mu\over {g^{1\over 7}_{s}}}\right)^{2}d\Omega^2_8 \cr &=&\alpha^{\prime}\left[-\left({{5}\over{2N^{1\over 7}}}\right)^2u^2dt^2+ \left({{2N^{1\over 7}}\over{5}}\right)^2{{du^2}\over{u^2}}+N^{2\over 7} d\Omega^2_8\right], \label{dualads} \end{eqnarray} where $u=\bar{r}/\alpha^{\prime}$. In the dual frame, in which the metric in the near horizon takes the above AdS form, the effective action (\ref{iibsg}) takes the following form: \begin{equation} S={1\over{16\pi G_{10}}}\int d^{10}x\sqrt{-G^{dual}} \left[e^{-{6\over 7}\phi}({\cal R}^{dual}+{{16}\over {49}} \partial_M\phi\partial^M\phi)-{1\over 4}e^{{6\over 7}\phi} F_{MN}F^{MN}\right]. \label{dualact} \end{equation} On the other hand, one can view the $D0$-brane solution (\ref{sgdzero}) as being magnetically charged under the Hodge-dual field strength to the field strength of the 1-form potential in the RR sector. Namely, the supergravity solution (\ref{sgdzero}) solves the equations of motion of the following effective action and is magnetically charged under the 7-form potential $A_{M_1\cdots M_7}$: \begin{equation} S^{\prime}={1\over{16\pi G_{10}}}\int d^{10}x\sqrt{-G^{str}} [e^{-2\phi}({\cal R}^{str}+2\partial_M\phi\partial^M\phi) -{1\over{2\cdot 8!}}F_{M_1\cdots M_8}F^{M_1\cdots M_8}], \label{hdgact} \end{equation} where $F_{M_1\cdots M_8}$ is the field strength of the 7-form potential $A_{M_1\cdots M_7}$. In the dual frame with the spacetime metric $G^{dual}_{MN}=e^{-{2\over 7} \phi}G^{str}_{MN}$, the action (\ref{hdgact}) takes the following form \cite{town3}: \begin{equation} S^{\prime}={1\over{16\pi G_{10}}}\int d^{10}x\sqrt{-G^{dual}} e^{-{6\over 7}\phi}\left[{\cal R}^{dual}+{{16}\over {49}} \partial_M\phi\partial^M\phi-{1\over {2\cdot 8!}} F_{M_1\cdots M_8}F^{M_1\cdots M_8}\right]. \label{dualhdgact} \end{equation} As in Ref. \cite{town3}, if one properly includes the factor involving $N$ in the Weyl transformation of the metric, i.e. $G^{str}_{MN}\to G^{dual}_{MN}=(Ne^{\phi})^{-{2\over 7}}G^{str}_{MN}$, then the new radial coordinate $u$ (with which the near horizon metric takes the AdS form in the horospherical coordinates) is related to $U$ as $u={2\over 5}{{U^{5/2}}\over{g_{YM}N^{1/2}}}$. This is reminiscence of the holographic UV/IR connection between the bulk and the boundary theory \cite{suss3,polc1}, if one identifies $u$ with the energy scale $E$ of the boundary theory \cite{town3}. In this case, the effective action in the dual frame takes the following form \cite{town3}: \begin{equation} S^{\prime}={{N^2}\over{16\pi G_{10}}}\int d^{10}x\sqrt{-G^{dual}} (Ne^{\phi})^{-{6\over 7}}\left[{\cal R}^{dual}+{{16}\over {49}} \partial_M\phi\partial^M\phi-{1\over {2\cdot 8!}}{1\over{N^2}} F_{M_1\cdots M_8}F^{M_1\cdots M_8}\right]. \label{orgdualact} \end{equation} The near horizon form of D0-brane solution to the associated equations of motion is: \begin{eqnarray} ds^2_{dual}&=&-{{u^2}\over{u^2_0}}dt^2+u^2_0{{du^2}\over{u^2}} +d\Omega^2_8, \cr e^{\phi}&=&{1\over N}(g^2_{YM}N)^{7\over{10}}\left({u\over{u_0}} \right)^{-{{21}\over{10}}}, \cr F_8&=&7N\,vol(S^8), \label{dualneard0} \end{eqnarray} where $u_0=2/5$. From this solution, one can see that there is the Freund-Rubin compactification \cite{freu} on $S^8$ of the $D=10$ action (\ref{orgdualact}) to the following 2-dimensional effective gauged supergravity action \cite{town3}: \begin{equation} S^{\prime}=N^2\int d^2x\sqrt{-g}(Ne^{\phi})^{-{6\over 7}} \left[{\cal R}+{{16}\over{49}}\partial_{\mu}\phi\partial^{\mu}\phi +{{63}\over 2}\right]. \label{2dact} \end{equation} The near-horizon D0-brane supergravity solution (\ref{dualneard0}) is reduced under this $S^8$ compactification to a domain solution \cite{pope1,pope2,town4}, which is supported by the cosmological term in the action (\ref{2dact}). \section{Generalized Conformal Mechanics of D0-branes} We consider a probe D0-brane with mass $m$ and charge $q$ moving in the near horizon geometry of the source $N$ D0-brane bound state. The gravitational field felt by the probe D0-brane in the string [dual] frame is given by Eqs. (\ref{sgdecoupl}) and (\ref{adsform}) [Eqs. (\ref{dualmet}) and (\ref{dualads})]. Here, once again, $r$ is the radial distance between the source and the probe D0-branes. The probe D0-brane moves with the 10-momentum $p=(p_M)=(p_t,p_1,...,p_9)$ and its time-component is the (static-gauge) Hamiltonian $H=-p_t$. The expression for the Hamiltonian of the probe D0-brane in the near horizon background of the source D0-brane can be obtained by solving the mass-shell constraint of the probe D0-brane. Unlike the case of Ref. \cite{kall1}, which studies a charged particle in the Einstein-Maxwell theory, the probe D0-brane satisfies the mass-shell constraint which is different from the ordinary constraint $0=(p-qA)^2+m^2= G^{MN}(p_M-qA_M)(p_N-qA_N)+m^2$. This is due to the non-trivial dilaton field that is present for the D0-brane solution. So, here we rederive the mass-shell condition for the case of D0-branes. The action for the probe D0-brane with the mass $m$ and the charge $q$ moving in the background of the source D0-brane has the following form: \begin{equation} S=\int d\tau\,L=\int d\tau\left(me^{-\phi}\sqrt{-G^{str}_{MN}\dot{x}^M \dot{x}^N}-q\dot{x}^MA_M\right), \label{dpartact} \end{equation} where $v^M=\dot{x}^M\equiv{{dx^M}\over{d\tau}}$ is the 10-velocity of the probe D0-brane. Note, since we choose the static gauge for the action, the worldline time $\tau$ and the target-space time $t$ are set to equal. Once again, $G^{str}_{MN}$ and $A_M$ are the fields produced by the source D0-brane. Note, in this action the metric $G^{str}_{MN}$ is in the string frame. The (generalized) momentum conjugate to $x^M(\tau)$ is \begin{equation} P_M=-{{\delta L}\over{\delta\dot{x}^M}}={{me^{-\phi}\dot{x}_M} \over\sqrt{-G^{str}_{MN}\dot{x}^M\dot{x}^N}}+qA_M. \label{conjmom} \end{equation} As usual, $p_M=m\dot{x}_M/\sqrt{-G^{str}_{MN}\dot{x}^M\dot{x}^N}$ is the ordinary 10-momentum of the D0-brane. From this, one obtains the following mass-shell constraint for the probe D0-brane in the string-frame background of the probe D0-brane: \begin{equation} G^{str\,MN}(P_M-qA_M)(P_N-qA_N)+m^2e^{-2\phi}=0. \label{massshell} \end{equation} To obtain the expression for the Hamiltonian $H=-P_t$ for the probe D0-brane mechanics, we solve this mass-shell constraint (\ref{massshell}). We consider the following general spherically symmetric $D=10$ metric Ansatz: \begin{eqnarray} G_{MN}dx^Mdx^N&=&-A(r)dt^2+B(r)dr^2+C(r)d\Omega^2_8 \cr &=&-A(r)dt^2+B(r)dr^2+C(r)[d\theta^2+\cos^2\theta d\psi^2_1+ \cos^2\theta\cos^2\psi_1d\psi^2_2 \cr & &+\cos^2\theta\cos^2\psi_1\cos^2\psi_2d\psi^2_3+\mu^2_id\phi^2_i], \label{sphansatz} \end{eqnarray} where \begin{equation} \mu_1=\sin\theta,\ \ \mu_2=\cos\theta\sin\psi_1,\ \ \mu_3=\cos\theta\cos\psi_1\sin\psi_2,\ \ \mu_4=\cos\theta\cos\psi_1\cos\psi_2\sin\psi_3. \label{mudefs} \end{equation} Then, the expression for the Hamiltonian $H$ takes the following form: \begin{equation} H={{P^2_r}\over{2f}}+{{g}\over{2f}}, \label{genhamilton} \end{equation} where \begin{eqnarray} f&\equiv&{1\over 2}A^{-{1\over 2}}Be^{-\phi}\left[\sqrt{m^2+e^{2\phi} (P^2_r+BC^{-1}\vec{L}^2)/B}+qA^{-{1\over 2}}A_te^{\phi}\right], \cr g&\equiv&Be^{-2\phi}\left[(m^2-q^2A^{-1}A^2_te^{2\phi})+C^{-1} e^{2\phi}\vec{L}^2\right]. \label{deffg} \end{eqnarray} Here, $A_t$ is the time component of the 1-form field $A_M$ of the near-horizon source D0-brane supergravity solution (\ref{sgdecoupl}) and $\vec{L}^2$ is the angular momentum operator of the probe D0-brane given by \begin{equation} \vec{L}^2=P^2_{\theta}+{{P^2_{\psi_1}}\over{\cos^2\theta}}+ {{P^2_{\psi_2}}\over{\cos^2\theta\cos^2\psi_1}}+ {{P^2_{\psi_3}}\over{\cos^2\theta\cos^2\psi_1\cos^2\psi_2}}+ \sum^4_{i=1}{{P^2_{\phi_i}}\over{\mu^2_i}}. \label{angop} \end{equation} First, we consider the probe D0-brane moving in the string-frame near horizon background of the source D0-brane. In the original work \cite{kall1} of the superconformal mechanics of a test charged particle in the near horizon geometry of the Reissner-Nordstr\"om black hole, it was necessary to redefine the radial coordinate so that the metric components $A$ and $B$ in Eq. (\ref{sphansatz}) satisfy the relation $A=B^2$ for the purpose of setting the factor $A^{-{1\over 2}}B$ in Eq. (\ref{genhamilton}) equal to 1. However, in the D0-brane case, as we will see, it is more convenient to work with the original form (\ref{sgdecoupl}) of near horizon metric, since the expression for the Hamiltonian becomes simpler in the original radial coordinate. By substituting the string-frame near horizon metric into the general formulae (\ref{genhamilton}) and (\ref{deffg}), one obtains the following Hamiltonian for the probe D0-brane in the string-frame near horizon background of the source D0-brane bound state: \begin{equation} H={{P^2_r}\over{2f}}+{{g}\over{2f}}, \label{stringharm} \end{equation} where $f$ and $g$ are given by \begin{eqnarray} f&=&{1\over 2}g^{-1}_s\left[\sqrt{m^2+g^2_s\left({\mu\over r}\right)^7 \left(P^2_r+{{\vec{L}^2}\over{r^2}}\right)}+q\right], \cr g&=&g^{-2}_s\left({\mu\over r}\right)^{-7} (m^2-q^2)+{{\vec{L}^2}\over {r^2}}, \label{gfforstring} \end{eqnarray} or in terms of the SYM theory variables \begin{equation} H={{P^2_U}\over{2f}}+{{g}\over{2f}}, \label{stringharmym} \end{equation} where $f$ and $g$ are given by \begin{eqnarray} f&=&{1\over 2}\alpha^{\prime\,{1\over 2}}g^{-2}_{YM} \left[\sqrt{m^2+{{\alpha^{\prime\,-1}g^6_{YM}N}\over{U^7}} \left(P^2_U+{{\vec{L}^2}\over{U^2}}\right)}+q\right], \cr g&=&\left({{\alpha^{\prime\,-1}g^6_{YM}N}\over{U^7}}\right)^{-1} (m^2-q^2)+{{\vec{L}^2}\over{U^2}}. \label{gfforstringym} \end{eqnarray} It is interesting that the term $A^{-1}A^2_te^{2\phi}$ in Eq. (\ref{deffg}) becomes 1 for the near horizon D0-brane solution (\ref{sgdecoupl}). So, the expressions for $f$ and $g$ becomes greatly simplified. And, in particular, in the extreme limit ($m-q\to 0$) of the probe D0-brane, the first term in $g$ drops out. This also generally holds for any dilatonic 0-brane supergravity solutions. Just as in the case of the test charged particle in the Reissner-Nordsr\"om black hole background, the mechanics of the probe D0-brane has the $SL(2,{\bf R})$ symmetry with the following generators: \begin{equation} H={{P^2_r}\over{2f}}+{g\over{2f}},\ \ \ \ K=-{1\over 2}fr^2,\ \ \ \ D={1\over 2}rP_r, \label{sl2rgen} \end{equation} where the Hamiltonian $H$ generates the time translation, $K$ generates the special conformal transformation and $D$ generates the scale transformation or the dilatation. These generators satisfy the following $SL(2,{\bf R})$ algebra: \begin{equation} [D,H]=H,\ \ \ \ [D,H]=-K,\ \ \ \ [H,K]=2D. \label{sl2ralg} \end{equation} This is the D0-brane generalization of conformal quantum mechanics studied in Refs. \cite{scm1,scm2,kall1}. In fact, the near horizon solution (\ref{sgdecoupl}), when uplifted as a solution of the 11-dimensional gravity (i.e. the 11-dimensional plane wave solution), is invariant under the $SU(1,1)\cong SL(2,{\bf R})$ isometry \footnote{At the 10-dimensional level, such $SL(2,{\bf R})$ transformations act on the near-horizon supergravity D0-brane solution in such a way that the constant $Q$ in the harmonic function $H$ transforms as if it is a ``field'' on the worldvolume and then is set to a constant after the transformations \cite{jevi2}.} generated by the scale transformation $\delta_D$, the special coordinate transformation $\delta_K$ and the time translation $\delta_H$ \cite{jevi1}. Under the time translation, $\delta_Ht=1$, $\delta_HU=0$ and $\delta_Hg_s=0$. Under the special coordinate transformation, $\delta_Kt=-(t^2+k{{g^2_{YM}}\over{U^5}})$, $\delta_KU=2tU$ and $\delta_Kg_s=6tg_s$. Finally, under the dilatation, $\delta_Dt=-t$, $\delta_DU=U$ and $\delta_Dg_s=3g_s$. And $A_t$ transforms as a conformal field of dimension 1. These infinitesimal transformations satisfy the following $SL(2,{\bf R})$ algebra just like the symmetry generators (\ref{sl2rgen}) of the probe D0-brane mechanics: \begin{equation} [\delta_D,\delta_H]=\delta_H,\ \ \ \ [\delta_D,\delta_K]=-\delta_K,\ \ \ \ [\delta_H,\delta_K]=2\delta_D. \label{sl2r11d} \end{equation} Note, the string coupling $g_s$ changes under the dilatation and the special coordinate transformation, and especially $g_s$ becomes time dependent after the special coordinate transformation is applied. Thereby, this $SL(2,{\bf R})$ isometry of the near horizon geometry does not extend to a conformal symmetry of the complete supergravity solution. The corresponding boundary theory, i.e. $0+1$ dimensional SYM matrix quantum mechanics, has the same $SL(2,{\bf R})$ symmetry, but the string coupling $g_s$ also transforms under this symmetry, unlike the case of D3-brane. However, as noted in Ref. \cite{jevi1}, since the dilaton coupling $g_s$ is related to the matrix model coupling constant, one can still think of `generalized' $SL(2,{\bf R})$ conformal symmetry in which the string coupling is now regarded as a part of background fields that transform under the symmetry. The `generalized' conformal symmetry therefore transforms a matrix model at one value of the coupling constant to another. Note, as pointed out in the previous section, D0-brane in the dual frame has description in terms of domain-wall solution after the compactification on $S^8$. So, this is also related to the fact that in the domain-wall/QFT correspondence the choice of horosphere (the hypersurface of constant $u$) for the Minkowski vacuum corresponds to a choice of coupling constant of a non-conformal QFT \cite{town3}. In the non-conformal case, the interpolation between the AdS Killing horizon (in the dual frame) and its boundary therefore corresponds to an interpolation between strong and weak coupling, or vice versa. The extreme limit of the probe D0-brane ($(m-q)\to 0$) can be interpreted as the M theory in the LCF, since in this case both the source and the probe D0-branes are in the BPS limit. In this case, the functions $f$ and $g$ are simplified to \begin{equation} f={1\over 2}g^{-1}_s\left[\sqrt{m^2+g^2_s\left({\mu\over r}\right)^7 \left(P^2_r+{{\vec{L}^2}\over{r^2}}\right)}+m\right],\ \ \ \ g={{\vec{L}^2}\over {r^2}}, \label{extfgstr} \end{equation} or in terms of the SYM theory variables \begin{equation} f={1\over 2}\alpha^{\prime\,{1\over 2}}g^{-2}_{YM} \left[\sqrt{m^2+{{\alpha^{\prime\,-1}g^6_{YM}N}\over{U^7}} \left(P^2_U+{{\vec{L}^2}\over{U^2}}\right)}+m\right],\ \ \ \ g={{\vec{L}^2}\over{U^2}}. \label{extfgstrym} \end{equation} Unlike the case of a charged test particle in the near horizon Reissner-Nordsr\"om black hole background studied in Ref. \cite{kall1}, we do not let the mass of the source D0-brane bound state go to infinity, since the number $N$ of the D0-branes in the M theory on the LCF is kept finite. The fact that we are considering the near horizon geometry of the D0-brane supergravity solution means that we are in the LCF of M theory, since the near horizon geometry of the D0-brane supergravity solution is also the null reduction of the M wave supergravity solution \cite{tsey1}, which is interpreted as M theory in the LCF. Taking $N$ to infinity, i.e. infinitely massive source D0-brane, corresponds to M theory in the IMF. Note, the IMF is defined as the limit in which $N$ and $R_{11}$ go to infinity such that the momentum $P_{11}=N/R_{11}$ also goes to infinity. Since $P_{11}\sim g^{-7/2}_{YM}N^{1/4}U^{21/4}$ and $R_{11}\sim g^{7/2}_{YM} N^{3/4}U^{-21/4}$, one can take both $P_{11}$ and $R_{11}$ to infinity while taking $N\to\infty$, if $U$ is in the range of $g^{2/3}_{YM}N^{-1/21}\ll U\ll g^{2/3}_{YM}N^{1/7}$. However, this range of $U$ is beyond the range of validity (\ref{rcnstrnt}) of the near horizon solution (\ref{sgdecoupl}). In other words, one cannot go to the IMF of M theory while keeping the parameters within the validity of the near horizon solution (\ref{sgdecoupl}). Anyway, the IMF means decompactification ($R_{11}\to\infty$) to the eleven dimensions. So, we should rather consider the M wave solution in the case $N\to\infty$. By expanding this Hamiltonian for an extreme $D0$-brane probe, one obtains a Hamiltonian of the form which is the sum of the non-relativistic kinetic term for the probe and the velocity dependent potential given by the sum of terms of the form $\sim {{v^{2n+2}}\over{r^{7n}}}$ ($n\in{\bf Z}^+$). This should reproduce the (M)atrix theory calculation of graviton-graviton scattering and its supergravity calculation reproduction of Ref. \cite{tsey1}. This is because the above action (\ref{dpartact}) is just the reduction of the action for the probe in the background of $D=11$ plane wave (describing the motion of the probe graviton in the background of the moving heavy source graviton) on a light-like circle to 10 dimensions, i.e. M theory in the LCF. Note also that since M theory in the LCF has ``Galilean'' invariance (or is described by non-relativistic quantum mechanics) in the transverse space, the above Hamiltonian for an extreme probe D0-brane has the non-relativistic structure. These above arguments, together with the fact that the generalized conformal mechanics of probe D0-brane and the boundary $SU(N)$ SYM theory (i.e. the (M)atrix model of M theory in the LCF) satisfy the same $SL(2,{\bf R})$ symmetry, implies the equivalence between the (M)atrix model with finite $N$ and the generalized conformal mechanics of D0-brane. Next, we consider the probe D0-brane moving in the background of the dual-frame metric of the source D0-brane bound state. Since the dual frame can be considered as a ``holographic frame'' describing supergravity probes \cite{town3}, it is worthwhile to consider the case of the dual frame. In this frame, the near horizon metric takes the AdS$_2\times S^8$ form. So, one should expect that the `generalized' conformal mechanics with the $SL(2,{\bf R})$ symmetry can also be realized in the dual frame. The action for the probe D0-brane in the dual frame background of the source D0-brane is given by (\ref{dpartact}) with the string-frame metric $G^{str}_{MN}$ replaced by the dual-frame metric $G^{dual}_{MN}$ through the relation $G^{str}_{MN}=e^{{2\over 7}\phi}G^{dual}_{MN}$. So, for the probe D0-brane in the dual frame, the dilaton factor $e^{-\phi}$ in Eqs. (\ref{dpartact}) and (\ref{conjmom}) [the dilaton factor $e^{-2\phi}$ in Eqs. (\ref{massshell}) and (\ref{deffg})] is replaced by $e^{-{6\over 7} \phi}$ [$e^{-{{12}\over 7}\phi}$]. Substituting the dual-frame near horizon solution (\ref{dualmet}) into this general expression for the Hamiltonian corresponding to the dual frame, one finds that the Hamiltonian in the dual frame has the same form as the string-frame Hamiltonian (\ref{stringharm}) with the same $f$ and $g$ (\ref{gfforstring}). In fact, in general the Hamiltonian describing probe D0-brane is independent of the near horizon spacetime frame of the source D0-brane. So, the dynamics of the probe D0-brane in the dual frame also has the $SL(2,{\bf R})$ symmetry with the same symmetry generators (\ref{sl2rgen}) and algebra (\ref{sl2ralg}) as the string-frame case. As pointed out in the previous section, the source D0-brane solution also has the (Hodge) dual description in terms of the 8-form field strength, whose magnetic charge now is carried by the source D0-brane. In this case, by compactifying the D0-brane solution on $S^8$ one obtains a domain wall solution in $1+1$ dimensions \cite{town4,town3}. This domain wall solution solves the equations of motion of a $D=2$ $SO(9)$ gauged maximal supergravity theory, which is an $S^8$ compactification of the type IIA supergravity. This $SO(9)$, which is the largest subgroup of the $SO(16)$ and the isometry group of $S^8$ upon which the D0-brane is compactified, is also the R-symmetry group of the corresponding boundary $D=1$ QFT. In general, the R-symmetry of the supersymmetric QFT on the domain wall worldvolume matches the gauge group, which is the isometry group of the compactification manifold, of the equivalent gauged supergravity \cite{town3}. Putting together all the above facts, namely ($i$) generalized conformal mechanics of the probe D0-brane in the string-frame near horizon background of the source D0-brane is related to the M theory in the LCF, ($ii$) the generalized conformal mechanics of the probe D0-branes in the string and the dual frames of the near horizon source D0-branes are described by the same Hamiltonian, ($iii$) upon the dimensional reduction on $S^8$ the bulk theory of the source D0-brane in the dual frame in the Hodge-dual description (in terms of the 8-form field strength) is the $D=2$ $SO(9)$ gauged maximal supergravity theory, one arrives at the speculation that the M theory in the LCF is related to the $D=2$ $SO(9)$ gauged maximal supergravity theory (i.e. a $D=2$ Kaluza-Klein supergravity theory with domain wall vacuum). \vskip2.mm
\section{Introduction} The quantization of superstrings in non-trivial backgrounds has become an urgent problem since it has been realized that there is a direct relationship to the strong coupling dynamics of gauge field theories. In particular, one is often interested in conformal field theories which are related to spacetimes which contain an anti-deSitter factor.\cite{Juan} For examples with $AdS_3$ and backgrounds of NSNS fields, several cases have been worked out\cite{GKS,EFGT,KLL,BORT,newKS} for finite $N$. The more interesting case with RR backgrounds, corresponding to configurations of D-branes, are much more difficult. Work on $AdS_3$ has been presented in Ref. \cite{BVW}, while $AdS_5\times S^5$ has been discussed in the Green-Schwarz formalism\cite{Metsaev:1998it,renata}. Ref. \cite{ARMR} discusses the latter in the context of an expansion around flat spacetime. In this paper, we consider RR backgrounds in the covariant RNS formalism. We begin with the theory in flat spacetime and demonstrate, that in the large $N$ limit, string perturbation theory organizes itself into a $\sigma$-model which is exactly conformal. This procedure may be thought of as summing of worldsheet holes, with closed string loops suppressed because of large $N$. We check conformal invariance of the string path integral explicitly at leading non-trivial order, an application of the Fischler-Susskind mechanism.\cite{WFLS} The $\sigma$-model is of a non-standard type because of the RR background, but can be written formally in a way which is useful for perturbative calculations. The case of $AdS_5\times S^5$ is of particular interest. This model has a great deal of symmetry and it is possible that a free conformal field theory description exists, perhaps similar to the NS backgrounds with $AdS_3$. There are a number of oddities, centering around the properties of superconformal ghosts, as well as the properties of the spacetime supersymmetry algebra, which remain elusive.\cite{BLL} The present construction however is certainly not such an exact description, and must be discussed in the context of a weak curvature expansion. The analysis may be applied to other cases as well, such as the $D1-D5$ system, but we confine our attention here to $AdS_5\times S^5$. The paper is organized as follows. In section 2, we discuss general properties of string perturbation theory and demonstrate that in the large $N$ limit, the leading effect corresponds to a sum over worldsheet boundaries. We argue that these effects can be accounted for by a formal $\sigma$-model which includes the RR background. In order to define the $\sigma$-model it is necessary to introduce an operator which acts like the square-root of the picture-changing operator, $P_{1/2}$. In section 3, we discuss the conformal invariance of this $\sigma$-model at lowest order in $\alpha'$ and demonstrate that the Fischler-Susskind mechanism properly resums the background geometry. In section 4, we discuss some simple aspects of the spectrum of excitations around this background, and demonstrate that at a given level, R and NS states mix. In section 5, we consider the realization of spacetime symmetry charges in terms of worldsheet currents and comment on the supersymmetry algebra. Our analysis here is limited by our incomplete knowledge of the properties of $P_{1/2}$. Further comments and speculations are reserved for the final section. \section{Background of D3-branes: Summing over Boundaries} The basic idea here for computing the partition function is familiar from field theory: consider a scalar field theory, with a background field turned on. The one-loop determinant may be thought of as being built up by summing over multiple insertions of the background. \centerline{\littlefig{485oneloopeff.EPSF}{3}} Indeed, in string theory in NSNS backgrounds, it is well known that summing over multiple insertions of vertex operators simply exponentiates those vertex operators, resulting in a $\sigma$-model. Here, we repeat this analysis in the case of interest, with RR-backgrounds, and we will discover in exactly what sense the RR vertex operator exponentiates. We will find that it is possible to write a formal expression (involving one term) that quantifies the effect of the RR-background. This is necessarily formal, because it involves an operator which changes ghost number in a non-standard way. The alternative is to remove this operator, but at the expense of introducing an infinite number of non-local operators into the ``$\sigma$-model''. The utility of the formal expression is that it automatically generates the correct combinatorics, and has a clear interpretation for any non-zero S-matrix element. Let us begin by discussing the classical geometry exterior to a collection of $N$ D3-branes. The background is given by \beql{bgmetric} ds^2=f(r)^{1/2}dx_\perp^2+f(r)^{-1/2}dx_\parallel^2 \end{equation} where $f(r)=1+R_s^4/r^4$, $r$ being the radial perpendicular coordinate. The radius of curvature satisfies $R_s^4=4\pi gN\alpha'^2$. In addition, there is a self-dual RR five-form fieldstrength, with total flux $N$, and the dilaton is constant. One notes that for large $gN$, the curvature of spacetime is everywhere small, and therefore solving the equations of motion at first order in $\alpha'$ is a very good approximation. Furthermore, it has been argued from spacetime considerations that the near-horizon geometry is exact to all orders in $\alpha'$ and $g$. Our first task is to understand these results directly from string perturbation theory, where we integrate open string loops to all orders. In fact, the parameter that controls this perturbative expansion is $R_s/r$, so we will consider very large $r$: we are expanding around flat spacetime. Furthermore, for $R_s^2>>\alpha'$, it is possible to neglect the effects of higher string modes, even in the near-horizon limit. In principle, since we are including open string loops to all orders, it would seem that we also need to include closed string loops. However, closed string loops are suppressed in large $N$: they contribute at order $g^2\sim (gN)^2\times {1\over N^2}$, and can be safely ignored. Open string loops contribute at order $gN$ since the worldsheet boundary can be placed at any of the $N$ D-branes. To begin, let us consider the effects of a single boundary. Polchinski\cite{RRPol} computed the tadpoles of the graviton and RR 4-form potential in the presence of a D-brane to determine that a single D-brane carried one unit of RR-charge. The same calculation can be used to deduce that the gravitational field at a distance $r$ from a D3-brane has strength $\tilde G(X)=R_s^4/2r^4$. In this calculation, the information about polarizations is lost. This information may be recovered by a consideration of scattering of gravitons off of a D-brane.\cite{Klebanov:1996ni,Gubser:1996wt} The result is that the polarization manifestly preserves $SO(6)\times SO(3,1)$, \begin{equation} g_{\mu\nu}(X)={R_S^4\over 2r^4} t_{\mu\nu} \equiv{R_S^4\over 2r^4}\left( \delta^\perp_{\mu\nu}-\delta^\parallel_{\mu\nu}\right) \end{equation} This is the leading term in an expansion of the supergravity metric \eq{bgmetric} and we will see that it is consistent from the point of view of the string path integral as well. If we are at large $r$, we can think of this as a gravitational fluctuation, with (on-shell) vertex operator in the (0,0) picture\footnote{Our normalizations are $\langle X(z)X(z')\rangle=-{\alpha'\over2}\ln|z-z'|^2$ and $\langle\psi(z)\psi(z')\rangle=1/(z-z')$. Furthermore, $G(X)={\tilde G(X)\over 4\pi}$ (the $4\pi$ is introduced to account for the differing normalizations of metric and vertex operator.)} \begin{eqnarray}\label{eq:gravitonvop} \delta V_g &=& -{2\over\alpha'}t_{\mu\nu} \left(-G(X)\pa X^\mu\bar\pa X^\nu +{\alpha'\over 2}(\psi\cdot\partial G(X))\psi^\mu\bar\pa X^\nu\right.\nonumber\\ & &\left.+{\alpha'\over 2}\pa X^\mu(\tilde\psi\cdot\partial G(X))\tilde\psi^\nu -\left({\alpha'\over 2}\right)^2 \psi^\lambda\psi^\mu\tilde\psi^\rho\tilde\psi^\nu \partial_\lambda\partial_\rho G(X)\right) \end{eqnarray} We will find that it is necessary to modify this vertex operator by contact terms; we will see that this is equivalent to going to $N=1$ superspace. The RR vertex operator in the $(-{1\over2},-{1\over2})$-picture takes the form\footnote{$C$ is charge conjugation, $\Gamma_{-1}$ is 10-dimensional chirality, and $\Gamma_{\perp(\parallel)}$ is chirality in the 6(4)-dimensional subspace.} \beql{RRvop} \delta V_{RR} =\tilde S^T C\Gamma^{\mu_1\ldots\mu_5}\left( {1+\Gamma_{-1}\over 2}\right) S H_{\mu_1\ldots\mu_5}e^{-\phi/2}e^{-\bar\phi/2} \end{equation} where \begin{equation} \Gamma^{\mu_1\ldots\mu_5}H_{\mu_1\ldots\mu_5} =\Gamma_\parallel\Gamma^\alpha\partial_\alpha G(X) \end{equation} The normalization of $V_{RR}$ may be determined by factorization. The first step is to check under which conditions the operator \eq{gravitonvop} is well-defined; instead of defining it to be normal-ordered, we require that the singularities are absent. There are both quadratic divergences (contact terms)\cite{MGNS} and logarithmic divergences to consider. We will confine our attention to the $\pa X\bar\pa X$ part of the graviton vertex operator. The other terms follow from worldsheet supersymmetry. The prefactor $G(X)$ will be dealt with by expansion around a fixed configuration: $G(X)=\langle G(X)\rangle+X^\mu\langle \partial_\mu G(X)\rangle+\ldots$. Point-splitting $\delta V_g$ (eq. \eq{gravitonvop}), we find \begin{eqnarray}\label{eq:selfcontract} \int d^2z\ \delta V_g=\int d^2z\ :\delta V_g: + {\rm tr\ } t\ 2\pi\delta^{(2)}(\epsilon)\int d^2z :G(X):\\ -{1\over 2}\ln |\epsilon|^2\ \int d^2z :\Box G(X) t_{\mu\nu}\pa X^\mu\bar\pa X^\nu: +\ldots\nonumber\end{eqnarray} In the ellipsis are the fermion terms. There are higher order terms in $\alpha'$, but they are all proportional to (derivatives of) $\Box G$. Now, because $G(X)$ is a harmonic function, $\Box G(X)$ vanishes, at least away from the branes. Thus $\delta V_g$ is defined up to a contact term. This contact term may be removed by modifying $\delta V_g$ by terms which vanish on the (tree-level) equations of motion. In fact, these modifications are equivalent to introducing $N=1$ superspace \begin{equation} {\bf X}=X+i\theta\psi+i\bar\theta\tilde\psi+\theta\bar\theta F \end{equation} whereupon the vertex operator is written: \beql{newgop} \delta V_g=t_{\mu\nu}\int d^2z\int d^2\theta\ G({\bf X})D{\bf X}^\mu\bar D{\bf X}^\nu. \end{equation} The new terms in the vertex operator which are relevant for the contact terms are \begin{equation} Gt_{\mu\nu}(\psi^\mu\bar\partial \psi^\nu+\tilde\psi^\nu\partial\tilde\psi^\mu+F^\mu F^\nu) \end{equation} The fermionic terms in the vertex operator are just the appropriate covariantizations \begin{equation} \psi^\mu(G\bar\partial \delta^\nu_\lambda +\bar\partial X^\nu \partial_\lambda G)\psi^\lambda, \end{equation} etc. These give additional contractions which cancel the contact term. Thus we have shown that the vertex $\delta V_g$ is well-defined. The RR vertex operator requires no such treatment, as there are no contact terms;\footnote{This assumes the spin field $S^\alpha$ is normal-ordered.} it is regular, as long as $\Box G=0$. With these operators in hand, we are prepared to consider string worldsheets with multiple insertions. We will find that there are additional logarithmic short-distance singularities, which can be canceled by the Fischler-Susskind mechanism, namely by modifying the function $G$. This calculation appears in Section 3. Let us consider the scattering of string states off of the D-branes at low momentum transfer. The momentum transfer is the Fourier transform of the coordinate $r$. We want to show that the contributions to this scattering are dominated by the exchange of massless states. The boundary on the D-brane may then be replaced in this factorization limit, by an insertion of a massless background vertex. Indeed, this is what we would expect from looking at a Born-Oppenheimer approximation to the scattering. In essence, if each boundary has this property, then the relevant limit for calculating amplitudes involves long thin tubes attached to the brane, so each boundary is effectively shrunk to zero size, and can be replaced by a local insertion of a massless vertex operator, which depends on $r$. This dependence comes about as each long thin tube extends from where the interaction takes place to the D-brane, and hence involves the Green's function $\sim {1\over r^4}$ for the massless state in consideration. The fact that the amplitude factorizes into a long thin tube for each boundary is a non-trivial statement, as one might imagine that there might be hard momentum flowing from one hole to another, thus making the boundaries big. We will argue that these effects cancel each other by supersymmetry. Indeed, in a low momentum scattering, the amplitude of external states factorizes onto a one point function of a vertex operator that is almost on-shell and massless, on a Riemann surface with $ m>1$ holes. As we have more than one hole in the surface, the diagram without insertion is a self energy diagram of the brane configuration, and by supersymmetry vanishes; addition of the massless state insertion to this diagram is zero on-shell, as it probes the self-energy of the brane externally. The argument breaks down for $m=1$, as this probes the 'tree' level contribution to the mass of the brane, and this is the D-brane tension. Thus, the relevant limit of moduli space that contributes to the amplitude involves long thin tubes for all the insertions, and hence we get only effects from massless vertex operators in the full calculation. Summing over this restricted moduli space of zero size holes integrates these vertex operator insertions over a Riemann surface with fixed complex structure. We sum over all inequivalent surfaces, and since the boundaries are all identical, we should divide out by the permutation symmetry to avoid overcounting. The sum over boundaries generates the series \begin{equation} 1+\sum_{m=1}^\infty \frac 1{m!}\left(\int_\Sigma V_B \right)^m = \exp{\int_\Sigma d^2 z V_B} \end{equation} with $V_B$ a massless string state generated by the boundary. That is, the sum over boundaries generates an expression which has a sigma model interpretation where we modify the action by \beql{genexp} S\to S+\int_\Sigma d^2 z V_B(r) \end{equation} where $r$ is the impact parameter quantum field on the worldsheet. The sum over spin structures implies that $V_B$ is given by the combined graviton and self dual RR tadpole which we extracted in eqs. \eq{newgop},\eq{RRvop} above. From the point of view of string perturbation theory around flat spacetime, this corresponds to a partial resummation. As far as the metric is concerned, the exponentiation is standard. For the RR background, as discussed, the combinatorics are right for exponentiation, but the superconformal ghost factors make this problematic. A $\sigma$-model for these backgrounds may be written, formally, in several ways. The basic issue is that we wish to write an expression with zero ghost charge. In an S-matrix element, it would be convenient to take the external states in a fixed picture, and the insertion of eq. \eq{genexp} implies that each term in $V_B$ should be in the same picture. There may exist a field redefinition which mixes ghosts and matter fields similar to Ref. \cite{Berkovits:1994wr,BVW} that would sidestep this problem but instead we propose to formally write the $\sigma$-model as \begin{equation} S_{RR}\sim \int P_{1/2}\bar P_{1/2} V_{(-1/2,-1/2)} \end{equation} where $P_{1/2}$ ($\bar P_{1/2}$) increases the L(R)-ghost charge by $1/2$ and satisfies $P_{1/2}^2=P_{+1}$ ($\bar P_{1/2}^2=\bar P_{+1}$). Such operators are clearly not well-defined. However, this expression has two important properties: {\it it generates the correct combinatorics, and is unambiguous for any non-zero S-matrix element}. The basic point is that in an expansion of this exponential, the RR background vertex only contributes in pairs (on any topology). Using this $\sigma$-model, one can systematically compute perturbative corrections. For a given on-shell S-matrix element on a given topology, one introduces the correct background ghost charge explicitly, then takes each vertex operator, and each background vertex in the $(0,0)$ picture, defined as above. Further issues involving the properties of $P_{1/2}$ may be found in Sections 5,6. An alternative procedure would be the following. Since the one-point function of $V_{RR}$ is zero on any topology, one could attempt to write non-local expressions of the form \begin{equation} S_{RR}\sim \int V_{(+1/2,+1/2)}\int V_{(-1/2,-1/2)}+\ldots \end{equation} but the combinatorics of such an action are not correct: one needs to correct this at order $V^4$ and so on. An action with an infinite number of non-local terms is certainly not a terribly convenient representation. \section{Fischler-Susskind and the $\sigma$-model} Having discussed the general form of string perturbation theory in this background, we now give an explicit calculation of the modifications to the background which follow from conformal invariance of the string path integral. The calculation is an application of the Fischler-Susskind mechanism for a pair of colliding worldsheet holes. Thus we consider two background vertex operators, defined above, and bring them close together. \smallskip \centerline{\littlefig{opprodFS.EPSF}{2}} \noindent The calculation of the short-distance singularity is somewhat involved but straightforward, and we present here only the result. The quadratic divergences (contact terms) cancel exactly, given the modified form of the vertices discussed above. For simplicity, we write here the contributions to only the $\pa X\bar\pa X$ part \begin{eqnarray} \int d^2z\ \delta V_g(z)\cdot\int d^2z'\ \delta V_g(z')=&&\nonumber\\ -4\pi\ln |\epsilon|^2\int d^2z \Bigg(\pa X^\mu\partial_\mu G\ \bar\pa X^\nu (t^2)_\nu^{\ \lambda}\partial_\lambda G\\ +\bar\pa X^\mu\partial_\mu G\ \pa X^\nu (t^2)_\nu^{\ \lambda}\partial_\lambda G \nonumber\\ -{1\over2}({\rm tr\ } t^2) \pa X^\mu\partial_\mu G\ \bar\pa X^\nu\partial_\nu G -\pa X^\mu t_\mu^{\ \rho}\partial_\rho G\ \bar\pa X^\nu t_\nu^{\ \lambda}\partial_\lambda G\nonumber\\ +(\partial G\cdot \partial G)(t^2)_{\mu\nu}\pa X^\mu\bar\pa X^\nu +G\Box G\ t_{\mu\nu}\pa X^\mu\bar\pa X^\nu\Bigg)\nonumber \end{eqnarray} Since $G(X)$ depends only on $r$, using the form of $t_{\mu\nu}$ discussed earlier this reduces to \begin{eqnarray} \int d^2z\ \delta V_g(z)\cdot\int d^2z'\ \delta V_g(z')=\nonumber\\ 4\pi\ln |\epsilon|^2\int d^2z \left(4\partial_\mu G\partial_\nu G-(\partial G\cdot \partial G)\eta_{\mu\nu}\right) \pa X^\mu\bar\pa X^\nu+\ldots \end{eqnarray} which is in the form of a graviton vertex operator. There is also a logarithmically divergent contribution of the same form from two background RR vertices. After some Dirac algebra, we find \begin{eqnarray} \int d^2z\ \delta V_{RR}(z)\cdot\int d^2z'\ \delta V_{RR}(z')=\nonumber\\ 4\pi\ln|\epsilon|^2\int d^2z \left(-4\partial_\mu G\partial_\nu G+2(\partial G\cdot \partial G)t_{\mu\nu}\right) \pa X^\mu\bar\pa X^\nu \end{eqnarray} To obtain this result, the normalization of the RR vertex operator is required, as given in \eq{RRvop}, and discussed in Section 5. The total logarithmic divergence in the graviton channel is then of the form \beql{totdiv} 4\pi\ln |\epsilon|^2\int d^2z (\partial G\cdot \partial G)\left( \delta^\perp_{\mu\nu}-3\delta^\parallel_{\mu\nu}\right) \pa X^\mu\bar\pa X^\nu+\ldots \end{equation} Note that the terms in $\partial r\bar\partial r$ have cancelled precisely between NS and R contributions.\footnote{Even if this were not true, a field redefinition can remove such terms.} In this sense, we have started in the ``right" coordinate system. The divergence \eq{totdiv} may be removed by modifying the graviton background through the addition of a term \begin{equation} -2\pi \left( {2\over \alpha'}\right) G^2(X) \left( \delta^\perp_{\mu\nu}-3\delta^\parallel_{\mu\nu}\right) \pa X^\mu\bar\pa X^\nu+\ldots \end{equation} where the ellipsis contains fermionic terms, determined by supersymmetry. These terms have a logarithmically divergent self-contraction which exactly cancels that of eq. \eq{totdiv}. To see this, one can use eq. \eq{selfcontract} with $G$ replaced by $-2\pi G^2$: since $G^2$ is not harmonic, there is a non-zero contribution from $\Box G^2$. At this order then, the graviton background has the form \begin{eqnarray} {1\over 2\pi\alpha'}\left\{ \left(\tilde G-{1\over2}\tilde G^2\right)\delta^\perp_{\mu\nu} +\left(-\tilde G+{3\over2}\tilde G^2\right)\delta^\parallel_{\mu\nu}\right\} \pa X^\mu\bar\pa X^\nu+\ldots\\ \simeq{1\over 2\pi\alpha'} \left\{ (f^{1/2}-1)\delta^\perp_{\mu\nu}+(f^{-1/2}-1)\delta^\parallel_{\mu\nu}\right\} \pa X^\mu\bar\pa X^\nu+\ldots \end{eqnarray} where $f=1+2\tilde G(r)$ is the harmonic function appearing in the supergravity metric. Thus we see that at lowest order, the Fischler-Susskind mechanism does build up properly the background geometry of the 3-branes. Similar calculations may be performed without difficulty to demonstrate the appropriate modification of the RR background. We believe that this works to all orders in ${R_s\over r}$; comments in this regard may be found in a later section of this paper. Certainly, this is not in disagreement with spacetime arguments that the near-horizon geometry is exact.\cite{MGTB,Kallosh:1998qs} We present the following worldsheet argument that this is exact to all orders in $R_s/r$. Since $\alpha'/R_s^2$ is small, it is useful to consider an expansion of the $\sigma$-model in powers of curvature. In order to simplify this expansion, one goes to normal coordinates around a point. In terms of this expansion in curvature, the Ramond background is a first order perturbation, and the curvature correction to the metric is second order. In contrast, for the coordinates we were using before, which are not normal coordinates, the first correction comes from the Christoffel symbols, which are set to zero locally on the normal coordinate system. This can be done for the full geometry, but for simplicity, we specialize now to the near horizon region. For $AdS_5\times S^5$, expanding around zero in the normal coordinates we have \begin{equation} ds^2 \sim \eta_{\mu\nu}dx^\mu dx^\nu + A x_{AdS}^2 dx^2_{AdS} -A x^2_{S_5} dx_{S_5}^2+\dots \end{equation} where $A$ is determined by the curvature of the space. Also \begin{equation} V_{RR} =P_{1/2}\bar P_{1/2} S^\alpha \tilde S^\beta h_{\alpha\beta} \end{equation} with $h_{\alpha\beta} \sim [C(\Gamma_{AdS} + \Gamma_{S_5})]_{\alpha\beta}$. The $\Gamma$ matrices used are products of the five gamma matrices tangent to $AdS_5$ and $S_5$ respectively. Again we consider the collision of two background vertices. The algebra is straightforward, and we proceed by cancelling the logarithmic corrections from the square of the Ramond background, and the curvature self-contraction. This produces the beta function for the graviton, and it reads \begin{equation} (R_{\mu\nu} + (H^2)_{\mu\nu})\log(|\epsilon|^2)=0 \end{equation} We have used the $P_{1/2}$ operators for the Ramond vertex. As we have two Ramond vertex operators and to leading order they are on-shell (they are constant), the factors of $P_{1/2}$ give us a picture-changing operator, and the result is unambiguous. We thus reproduce the supergravity equations of motion for the graviton. This computation generalizes to any weakly coupled string theory at small curvature in the presence of RR backgrounds and gives the supergravity equations of motion. This computation is one-loop in the $\sigma$-model and thus is correct to leading order in $\alpha'$. Again according to \cite{MGTB,Kallosh:1998qs}, this result is exact for the near-horizon geometry to all orders in $\alpha'$. \section{Spectrum: Mixing of RR and NSNS states} We have suggested in the above discussions that the $\sigma$-model may be consistently defined, and it is exactly conformal. The next step would be to discuss the spectrum of excitations around the background. Since the $\sigma$-model is known and exactly conformal, in principle one could write equations for vertex operators which make them of dimension $(1,1)$. The best we can do at present is to compute vertex operators for these excitations perturbatively in the normal coordinate expansion, starting with flat spacetime operators. Note that in the approximation in which we are working, the spacing of states is as shown in the accompanying figure. The gap between levels is of order $\sqrt{\alpha'}$ while the spacing between $S^5$ harmonics is of order $1/R_s$. \smallskip \centerline{\littlefig{spectrumlevels.EPSF}{3}} We expect this spectrum to be modified by the background. The leading effect is to mix states at the same level. Because of the RR background, the states which in flat space are RR and NSNS mix with one another, which means that there are no longer separately conserved number operators in the R- and NS-sectors. At most, we can expect that a linear combination is still a good quantum number, and thus organizes the spectrum, although it seems unlikely that even this is true at finite $N$. Let us consider the one-loop partition function. In flat spacetime, bosonic symmetries which act on states come in the NS sector, and thus spacetime symmetries may be thought of as acting on, for example, the NSNS part of the partition function and the RR part {\it separately}. Once we turn on the background, this can no longer be true: there will be spacetime symmetries in the RR sector and thus at best we can classify states as bosonic or fermionic (which cancel by supersymmetry). In this section, we simply demonstrate this mixing at lowest order in $\alpha'$. The mixing is of order one. In principle, one can systematically compute higher order corrections. The basic idea is again an application of the Fischler-Susskind mechanism. We consider some arbitrary $(1,1)$ operator ${\cal O}e^{ip\cdot X}$ inserted on the worldsheet. We compute the effects of the background perturbatively by considering the effects of background vertex operators. \smallskip \centerline{\littlefig{spectrumFS.EPSF}{2}} The leading effect will come from a single RR insertion \begin{equation} \delta V_{RR}=\int d^2z\ P_{1/2}\bar P_{1/2} (CH)_{\alpha\beta}\tilde S^\alpha S^\beta e^{-\phi/2}e^{-\bar\phi/2} \end{equation} The operator product of $\delta V_{RR}$ with the vertex ${\cal O}$ will be of the form \begin{equation} \int d^2z\ {1\over |z-z'|^2} (CH)_{\alpha\beta} \left[ \bar Q^\alpha Q^\beta {\cal O}e^{ip\cdot X}\right] \end{equation} This is logarithmically divergent, and converts NS to R and vice versa. If need be this can be brought back to, say, the $(0,0)$ or $(-{1\over2},-{1\over2})$-picture by applying the picture-changing operator.\footnote{Note that at lowest order, we can assume that the picture changing operator is unchanged from its flat-space value. Higher order corrections must take this consistently into account.} Thus the leading order effect is to mix NS with R: \begin{equation} \delta\pmatrix{{\cal O}_{NS}\cr {\cal O}_R}= \pmatrix{0&M\cr \tilde M& 0}\pmatrix{{\cal O}_{NS}\cr {\cal O}_R} \end{equation} where $M$ and $\tilde M$ are suitable (logarithmically divergent) mixings that may be computed from the operator products mentioned above (we consider an explicit example below.) In order to eliminate the divergence, we should modify the vertex operators, in two steps. First, we diagonalize \begin{equation} {\cal O}_D\sim \alpha {\cal O}_{NS}+\beta {\cal O}_R \end{equation} and then modify the operator ${\cal O}_D$ to eliminate the divergence. In the weak curvature limit, it is sufficient to consider \begin{equation} :{\cal O}_D\ e^{ik\cdot X}:\to :{\cal O}_De^{ik\cdot X}: e^{i\delta k\cdot X} \end{equation} Going beyond the weak curvature approximation is a difficult problem, equivalent to solving for the explicit form of the vertex operators in this background, and we do not attempt such an analysis here. The modification of the momenta accounts for the change in the mass of states in the background and is the only source of a logarithmic divergence within the vertex operator. In principle, this analysis may be carried out for any state in the string spectrum, although it is technically challenging. We confine the discussion to a demonstration of the mixing effect for components of the massless gravity multiplet. For these states, the analysis reduces to that of Ref. \cite{Kim:1985ez} for supergravity in $AdS_5\times S_5$. Begin then with massless states: \begin{eqnarray} {\cal O}_{NS}&=&\int d^2z\ h^{(NS)}_{\mu\nu}(-\pa X^\mu+ik\cdot\psi\psi^\mu) (-\bar\pa X^\nu+ik\cdot\tilde\psi\tilde\psi^\nu) e^{ik\cdot X}\\ {\cal O}_R&=&\int d^2z\ h^{(R)}_{\alpha\beta}\tilde S^\alpha S^\beta e^{-\phi/2}e^{-\bar\phi/2}e^{ik\cdot X} \end{eqnarray} We note that \begin{equation} V_B\cdot {\cal O}_{NS} =-\int d^2z'\ e^{-\phi/2}e^{-\bar\phi/2}e^{ik\cdot X}S^\alpha\tilde S^\beta\cdot 2\pi\ln\epsilon\ h^{(NS)}_{\mu\nu}k_\lambda k_\rho (C\Gamma^{\lambda\nu}H\Gamma^{\rho\mu})_{\alpha\beta}\nonumber \end{equation} and \begin{equation} V_B\cdot {\cal O}_{R} =\int d^2z\ 2\pi\ln\epsilon\ h^{(R)}_{\alpha\beta} (C\Gamma^\mu H\Gamma^\nu)^{\alpha\beta} \psi^\mu\tilde\psi^\nu e^{-\phi}e^{-\bar\phi} \end{equation} These results are equivalent to a mixing matrix of the form: \begin{equation} -2\pi\ln\epsilon\ \pmatrix{0&h^{(NS)}_{\mu\nu}k_\lambda k_\rho (C\Gamma^{\lambda\nu}H\Gamma^{\rho\mu})_{\alpha\beta}\cr -h^{(R)}_{\alpha\beta} (C\Gamma^\mu H\Gamma^\nu)^{\alpha\beta} &0} \end{equation} and agrees with the supergravity analysis of \cite{Kim:1985ez}. This mixing suggests that the spacetime symmetries will not respect the splitting between R and NS. \section{Spacetime Algebra Let us consider the realization of spacetime symmetries on the worldsheet and their algebra. In this discussion, there are some subtleties because of limited knowledge of $P_{1/2}$. However, modulo these subtleties, we are able to reproduce the correct spacetime supersymmetry in the curvature expansion around the near-horizon geometry. A fuller discussion will appear in a subsequent publication. We know that in the presence of D-branes, half of the supersymmetry is broken, but is recovered in the near-horizon region as a superconformal symmetry. Our first task is to recover this fact in string perturbation theory. It is natural to define supercharges \begin{eqnarray} q^\alpha&=&\oint dz\ S^\alpha e^{-\phi/2}\\ \bar q^\alpha&=&\oint d\bar z\ \tilde S^\alpha e^{-\bar\phi/2} \end{eqnarray} which certainly generate symmetries in flat spacetime. In a RR background, at lowest order in $\alpha'$, there is only one linear combination \beql{mixedsuperch} Q^\alpha\sim P_{1/2}q^\alpha+\Gamma_\parallel\bar P_{1/2}\bar q^\alpha \end{equation} retained. To see this, we simply need to require that the supersymmetry annihilates the background \beql{bgsusy} Q^\alpha\cdot V_{B}=0 \end{equation} This calculation in fact is another way to fix the normalization of the RR vertex operator. At lowest order, our lack of understanding of $P_{1/2}$ is irrelevant. At higher orders, we need to know $P_{1/2}$, and furthermore we expect that eq. \eq{mixedsuperch} is further modified by functions of $r$. Once this happens, there is no possible splitting between holomorphic and anti-holomorphic fields. This effect can be computed from worldsheet considerations (given $P_{1/2}$) by a further application of Fischler-Susskind. From the spacetime point of view, they are determined by the requirement that $Q$ be associated to a Killing spinor. We note the very important point here that the holomorphic and anti-holomorphic charges mix with one another in the RR background. There is nothing really exotic about this: given a conserved current $J$, we can write $\bar\partial J+\partial \bar J=0$, and the spacetime charge $T=\oint dz\ J+\oint d\bar z\ \bar J$ is well-defined, being independent of contour. Only in special cases are $J$ and $\bar J$ separately conserved. In the near-horizon limit, we expect that exactly this happens, and the other linear combination of spacetime supercharges $S^\alpha$, comes back. Indeed, in normal coordinates for the near-horizon case, we begin with 32 supercharges, and eq. \eq{bgsusy} does not eliminate any. Instead, \eq{bgsusy} defines the modifications to the supercharges in the presence of the background. We may write the Killing spinor equation in the form \beql{killing} \nabla Q+Q\cdot V_B=0 \end{equation} which is the more general form of \eq{bgsusy}. Consider solving this equation order by order in the curvature expansion: \begin{eqnarray} Q^\alpha&=&Q^\alpha_{(0)}+Q^\alpha_{(1)}+\ldots\\ \bar Q^\alpha&=&\bar Q^\alpha_{(0)}+\bar Q^\alpha_{(1)}+\ldots \end{eqnarray} The lowest order terms are the flat space expressions $Q_{(0)}=q$, $\bar Q_{(0)}=\bar q$. At next order, the RR background in $V_B$ gives a non-zero contribution to the second term in eq. \eq{killing} which is proportional to $\bar P_{1/2}\bar q^\alpha$. Since in the normal coordinate expansion, the Christoffel symbols vanish, \eq{killing} then implies that \begin{equation} Q_{(1)}\sim \oint d\bar z \bar P_{1/2} X\cdot \Gamma\tilde S \end{equation} with a similar expression for $\bar Q_{(1)}$. Thus we see that indeed the supercharge has both holomorphic and antiholomorphic contributions. This is precisely what we need to reproduce the spacetime algebra. With the expressions that we have given, one may check the following commutators are induced by the worldsheet representation: \begin{eqnarray} \{ Q^\alpha, \bar Q^\beta\} &\sim & \oint (\Gamma_{\mu\nu})^{\alpha\beta} (X^\mu\cdot \pa X^\nu-X^\mu\cdot\bar\pa X^\nu)+\ldots\\ \{ Q^\alpha, Q^\beta\} &\sim & \oint (\Gamma_\mu)^{\alpha\beta}\pa X^\mu+\ldots \end{eqnarray} In general, in order to compute such effects precisely, we need to understand in more detail the operator $P_{1/2}$. We believe that these obstacles can be overcome. \section{Conclusions and discussion} In this paper, we have analyzed the background produced by D3-branes. We have discussed how a $\sigma$-model description can be used. The background geometry is recovered by summing over boundaries of worldsheets, and systematically applying the Fischler-Susskind mechanism. Our results suggest to us that an exact string conformal field theory description may exist. In order to `derive' such a description from perturbation theory requires further understanding of the square root of the picture changing operator. Our results indicate that this is not such an unnatural object. With this operator we can construct a perturbation expansion which is independent of the zero mode of the bosonized superconformal ghost system. This indicates that picture-changing is still a property of these $\sigma$-models, but the detailed form is modified by the background. It is plausible that the approach of \cite{BVW} is related to these remarks. The case of $AdS_3$ discussed there is considerably different however. In particular, there are half as many supersymmetries as for $AdS_5$, and consequently more freedom to choose pictures, and as well the target space of the $\sigma$-model is a group manifold. It seems difficult to imagine that a free-field realization exists in that the mixing of holomorphic and anti-holomorphic currents would seem to invalidate the idea of a spectrum generating algebra. \medskip \noindent {\bf Acknowledgments:} We wish to thank F. Larsen for collaboration at an early stage of this work. Research supported in part by the United States Department of Energy grant DE-FG02-91ER40677 and an Outstanding Junior Investigator Award. \providecommand{\href}[2]{#2}\begingroup\raggedright
\section{Introduction} The theoretical stellar yields of nitrogen are rather uncertain. It has been found (Renzini \& Voli 1981; Marigo et al 1996,1998; van den Hoek \& Groenewegen 1997) that nitrogen production takes place in the asymptotic giant branch phase of the evolution of intermediate-mass stars, but the mass range of the nitrogen-producing stars and the predicted amount of freshly produced nitrogen depend on poorly known parameters. The theoretical yields of nitrogen by massive stars are essentially unknown. Therefore, an "empirical" approach (analysis of the relation of N/O with O/H) has been widely used to study of the origin of nitrogen (Edmunds \& Pagel 1978; Lequeux et al 1979; Matteucci \& Tosi 1985; Garnett 1990; Pilyugin 1992, 1993; Vila-Costas \& Edmunds 1993; Marconi et al 1994; among others). Pagel (1985) called attention to the large scatter in N/O at fixed O/H in low-metallicity dwarf galaxies. Given the time delay between the injection of nitrogen by intermediate-mass stars and that of oxygen by shorter lived massive stars (the time -- delay hypothesis: Edmunds \& Pagel 1978) and the hypothesis of self-enrichment of star formation regions (Kunth \& Sargent 1986), models for the chemical evolution of dwarf galaxies reproducing the observed scatter of N/O have been constructed (Garnett 1990, Pilyugin 1992, Marconi et al 1994). A significant part (if not all) of nitrogen was assumed in these models to be produced by intermediate-mass stars. The HII regions with high N/O abundance ratios have been considered to be in early stages of star formation (before self-enrichment in oxygen occured), reflecting the chemical composition of the whole galaxy. The HII regions in advanced stages of the star formation bursts would have small N/O ratios due to self-enrichment in oxygen. These HII regions would then reflect the local chemical composition, and the scatter in their N/O ratios would be caused by different degrees of temporal decrease of N/O ratio within the HII regions. During the interburst period the nitrogen is ejected by the intermediate-mass stars, and the matter ejected by massive and intermediate-mass stars is supposed to be well mixed throughout the whole galaxy. Some scatter in global N/O ratios among dwarf galaxies can be caused by enriched galactic winds (Pilyugin 1993, 1994; Marconi et al 1994). In this framework, the chemical composition of low-metallicity dwarf galaxies was expected to be characterised by high N/O ratios with a relatively small dispersion, and the large observed dispersion in the N/O ratios is supposed to be due to a temporary N/O decrease inside HII regions in which the chemical composition is determined spectroscopically. In contrast with previous measurements, Thuan et al (1995) and Izotov \& Thuan (1999) have found a remarkably small scatter in the N/O ratios of the HII regions in low-metallicity (with 12+logO/H $\leq$ 7.6) BCGs. Izotov \& Thuan concluded that these galaxies are presently undergoing their first burst of star formation, and that nitrogen in these galaxies is produced by massive stars only. Since the intermediate and low-mass stars certainly do not make an appreciable contribution to oxygen production, the N/O ratio corresponding to the ejecta of massive stars and observed in low-metallicity BCGs is a lower limit for the global N/O ratios. This conclusion is in conflict with the fact that the nitrogen to $\alpha$-element abundance ratios measured in some DLAs are well below than the typlical value observed in low-metallicity BCGs (Lu et al 1998, and references therein). [Since oxygen abundance measurements are not available for DLAs, $[N/S]$ or $[N/Si]$ ratios are considered instead of $[N/O]$. This is justified by the fact that there is no reason to believe that the relative abundances of O, S and Si which are all produced in Type II supernovae are different from solar in DLAs.] However, as suggested by Izotov \& Thuan (1999), the nitrogen to $\alpha$-element abundance ratios in DLAs can be significantly underestimated if the absorption lines originate in the HII instead of the HI gas. Nevertheless, the possibility of a truly low nitrogen abundances in some DLAs cannot be excluded without additional observations, and these DLAs with low nitrogen to $\alpha$-element abundance ratios do not confirm the idea that the N/O ratio in low-metallicity BCGs is a lower limit for the global N/O ratio in galaxies. The best way to find the lower limit of N/O ratio and hence the amount of nitrogen produced by massive stars would be to determine the N/O ratios in galactic halo stars. Unfortunately, at the present state their N/O ratios cannot be determined with a precision better than a factor 2 or 3. The only firm conclusion is that nitrogen has a strong primary component (Carbon et al 1987). The goal of this study is to test whether the constancy of the N/O ratios in low-metallicity BCGs and and the scatter of the N/Si ratios in DLAs are compatible with each other and with the existing ideas on the chemical evolution of galaxies. \section{Possible interpretation of nitrogen abundances in BCGs and DLAs} Here we will demonstrate that the constancy of the N/O ratios in low-metallicity BCGs and the scatter of the N/Si ratios in DLAs can be reconciled under the following three assumptions: {\it 1}) a significant part of nitrogen is produced by intermediate-mass stars, {\it 2}) star formation in BCGs and DLAs occurs in bursts separated by quiescent periods, {\it 3}) the previous star formation events are responsible for heavy element abundances observed in the HII regions of the BCGs. As it was discussed in Introduction, there are no crucial arguments in favor of or against assumption {\it 1}. Assumption {\it 2} - that star formation in BCGs and DLAs occurs in bursts separated by quiescent periods - is commonly accepted in the case of BCGs after it was initially suggested by Searle \& Sargent (1972). In order to reproduce the observed properties of low-metallicity BCGs, only a few (in some cases only one or two) star formation bursts during their life are required (Tosi 1994). Papaderos et al (1998) have found that the spectrophotometric properties of SBS 0335-052 can be accounted for by a stellar population not older than $\sim$ 100 Myr. The possibility of an underlying old (10 Gyr) stellar population with mass not exceeding $\sim$ 10 times that of young stellar population mass hawever cannot be definitely ruled out on the basis of the spectrophotometric properties. In other words, the time interval between possible previous and current star formation events can be as large as $\sim$ 10 Gyr. In the case of DLAs, assumption {\it 2} seems to be also acceptable, despite the fact that DLAs do not constitute an homogeneous class of galaxies but belong to the wide variety of morphological types of galaxies (Le Brun et al 1997). Star formation in a given region in galaxies of any morphological type seems to be episodic. The DLA observations sample the general interstellar medium at random times along random lines of sight and may or may not see a region where a star formation event occured in a recent past. Assumption {\it 3} - that previous star formation events are responsible for the observed heavy element abundances in BCGs - is equivalent to say that the element abundances of HII regions in BCGs are not yet polluted by the stars of the present star formation event, and that their abundances reflect the average N/O in the galaxy, which results from cumulative previous star formation. Martin (1996) has found that the current event of star formation in the most metal-poor known blue compact galaxy I Zw 18 started 15-27 Myr ago. The duration of current star formation burst in another extremely metal-poor blue compact galaxy SBS 0335-052 (Papaderos et al 1998) is also in excess of the lifetime of the most massive stars. Therefore, a selection effect in favor of observations of young HII regions in which the massive stars had not yet have time to explode as supernovae cannot be only reason why the HII regions are not observed as self-enriched. Massive stars in the current star formation burst have often had time to synthesize heavy elements and to eject them via stellar winds and supernova explosions into the surrounding interstellar gas. Kunth \& Sargent (1986) suggest that the heavy elements produced in this way initially mix into H II region only, i.e. the giant H II regions are self-enriched. However, it is possible that the nucleosynthetic products of massive stars are in high stages of ionization and do not make appreciable contribution to the element abundance as derived from optical spectra (Kobulnicky \& Skillman 1997, Kobulnicky 1999). Indeed, the oxygen abundance in SBS 0335-052 has been measured within the region of 3.6 kpc (Izotov et al 1997). There is a supershell of radius $\sim$ 380 pc. There is no difference in oxygen abundances inside and outside the supershell as it should be expected since $\sim$ 1500 supernovae are required to produce this supershell (Izotov et al 1997). Other star-forming galaxies, which are chemically homogeneous despite the presence of multiple massive star clusters, are reported by Kobulnicky \& Skillman (1998). This can be considered as evidence that the nucleosynthetic products of massive stars in giant HII regions are hidden from optical spectroscopic searches because they are predominantly found in a hot, highly -- ionized superbubble. It should be noted however that some fraction of supernova ejecta can mix with dense clouds changing their chemical composition. If such cloud survives and produces a subgroup of stars shortly, the star formation region will have sub-generations of stars with different chemical composition. This seems to be the case in the Orion star formation region (Cunha \& Lambert 1994, Pilyugin \& Edmunds 1996). With assumptions {\it 1} and {\it 2} the behaviour of the N/O ratio is described by models of type suggested by Garnett (1990), Pilyugin (1992), Marconi et al (1994). Figgure \ref{figure:8566f1} illustrates the time behaviour of N/O in the interstellar medium of a galaxy in the case when all the nitrogen is produced by intermediate-mass stars (see also Fig.7 in Pilyugin 1992). Since only low-metallicity galaxies are considered here, the adopted total astration level in the models is small (less than 0.05), and the nitrogen yield is assumed to be independent on metallicity. The nitrogen yields by stars in the mass interval 3$\div$4M$_{\odot}$ were taken from Renzini and Voli (1981). As can be seen on Fig.\ref{figure:8566f1}, the N/O ratio increases for about 1 Gyr after the star formation burst and then remains constant. This is due to the assumption that nitrogen is not produced by stars with masses less than $\sim$ 3$M_{\odot}$ (Renzini and Voli 1981). If stars with masses less than $\sim$ 3$M_{\odot}$ also make some contribution to the nitrogen production (Marigo et al 1996, van den Hoek and Groenewegen 1997), this does not change appeciably the picture because the amount of nitrogen ejected decreases strongly with decreasing stellar mass (Renzini and Voli 1981, Marigo et al 1996, van den Hoek and Groenewegen 1997). Therefore, low-metallicity systems with a large time interval (more than $\sim$ 1 Gyr) between successive star formation events will have close values of N/O ratios before the current star formation event (Fig.\ref{figure:8566f1}). \begin{figure}[thb] \vspace{7.0cm} \special {psfile=pilyugf1.ps hscale=55 vscale=55 hoffset=-20 voffset=200 angle=-90.0} \caption{\label{figure:8566f1} The $[N/O]$ as a function of time for two models with different star formation histories. The dashed line corresponds to the evolution of a system in which two star formation bursts of equal intensities occured 12 Gyr and 3 Gyr ago. The solid line corresponds to the evolution of a system in which two star formation bursts occured 12 Gyr and 5 Gyr ago, the second star formation event being 5 times stronger than the first one.} \end{figure} With this behaviour of the N/O ratio and taking into account assumption {\it 3}, the nitrogen abundances in BCGs and DLAs can be interpreted in the following way. The low-metallicity BCGs are systems with a small amount of old (with age $>$ 1 Gyr) underlying stellar population over which the current star formation burst is superposed; only the stars from the previous star formation event(s) are responsible for the observed chemical composition in the giant HII regions in these galaxies. The DLAs with low nitrogen to $\alpha$-element ratios correspond to systems probed less than 1 Gyr after the last local star formation event, but after a time sufficient for disappearance of the superbubble and mixing of the freshly produced heavy elements in the interstellar medium. Conversely, the DLAs with nitrogen to $\alpha$-element ratios close to that in low-metallicity BCGs correspond to systems in which the time interval after last star formation event is sufficiently large for intermediate-mass stars to have substantially enhanced the nitrogen to $\alpha$-element abundance ratios. \section{Discussion and conclusions} We have shown that the observed nitrogen abundances in both BCGs and DLAs can be reproduced if a significant part of nitrogen is produced by intermediate-mass stars. If one assumes instead that the nitrogen abundances measured in low-metallicity BCGs is produced by massive stars, the observational data for BCGs and DLAs are incompatible with each other. Can this be considered as a crucial argument in favor of dominant production of nitrogen by intermediate-mass stars? Since the low nitrogen to $\alpha$-element abundance ratios obtained in some DLAs is not beyond question, the possibility that nitrogen in low-metallicity BCGs is produced by massive stars cannot be excluded. The crucial argument in favor of dominant production of nitrogen by intermediate-mass stars would be the reliable proof of the existence of systems with low N/O ratios. Solid determinations of nitrogen abundances in damped Ly$\alpha$ absorbers can clarify this matter. The Zn/S abundance ratios in DLAs with low measured nitrogen to $\alpha$-element abundance ratios can also tell us something about the reality of low N/O ratios in these objects. The measured value of $[Zn/S]$ in DLA associated with QSO 0100-130 is close to solar ratio (Prochaska \& Wolfe 1998, Lu et al 1998), indicating that Type I supernovae have already contributed to the zink abundance. The measured $[N/S]$ in this DLA is close to the $[N/O]$ ratio in low-metallicity BCGs. If values of Zn/S in DLAs with low N/O ratios are close to the solar ratio, this will be a strong argument in favor of an underestimation of the nitrogen to $\alpha$-element abundance ratios. The time scale for nitrogen enrichment is shorter than the time scale for iron enrichment, and a system with high Fe/O ratio should not have a low N/O ratio. If values of $[Zn/S]$ in DLAs with measured low N/O ratios are close to $[Fe/O]$ ratios in the galactic halo stars, this can be considered as an argument in favor of genuinely low nitrogen to $\alpha$-element abundance ratios in these DLAs. Of course, the best way to find the amount of nitrogen produced by massive stars would be an undisputable determination of N/O ratios in galactic halo stars. In summary: If a significant part of nitrogen is produced by intermediate-mass stars, it is possible to reconcile the observational data for BCGs and DLAs. If nitrogen is mainly produced by massive stars, the observational data for BCGs and DLAs are incompatible with each other. Since the low nitrogen to $\alpha$-element abundance ratios obtained in some DLAs are not beyond question, the possibility that nitrogen measured in low-metallicity BCGs is produced by massive stars cannot however be completely excluded. \begin{acknowledgements} I would like to thank Drs. N.G.Guseva and Y.I.Izotov for fruitful discussions. I thank the referee, Prof. J.Lequeux, for helpful comments and suggestions which resulted in a better presentation of the work. This work was partly supported through INTAS grant 97-0033. This study has been done using the NASA's Astrophysical Data Service. \end{acknowledgements}
\section{Introduction} \label{Introduction} Some recent works have been devoted to the interaction of ultracold atoms with microwave cavities \cite{Scu96,Mey97,Lof97,Sch97,Ret98,Zha99}. These studies treated the interaction between an incident atom in an excited state and a cavity field containing $n$ photons, taking the quantum mechanical CM motion of the atom into account. This interaction leads to a new kind of induced emission intimately associated with the quantization of the CM motion, named the mazer action \cite{Scu96}. After interaction with the cavity, the atom can be found transmitted in the excited state or in the lower state, or reflected as well. It has been shown \cite{Scu96} that these event probability amplitudes (denoted respectively $T_{a,n}$, $T_{b,n+1}$, $R_{a,n}$ and $R_{b,n+1}$) are given by \begin{equation} \label{TRinExpr} \renewcommand{\arraystretch}{1.4} \begin{array}{cc} T_{a,n}=\frac{1}{2}( t_{+n}+t_{-n} ) \qquad & T_{b,n+1}=\frac{1}{2}( t_{+n}-t_{-n} ) \\ R_{a,n}=\frac{1}{2}( r_{+n}+r_{-n} ) \qquad & R_{b,n+1}=\frac{1}{2}( r_{+n}-r_{-n} ) \end{array} \renewcommand{\arraystretch}{1.0} \end{equation} where $t_{\pm n}$ and $r_{\pm n}$ are, respectively, the transmission and reflection amplitudes of the elementary scattering process of the atom incident upon the potential \begin{equation} V_{n}^{\pm}(x)=\pm \hbar g \sqrt{n+1}\:u(x) \end{equation} where $x$ is the atom traveling direction, $g$ is the atom~-- field coupling strength and $u(x)$ is the cavity field mode function. The induced emission probability of the atom interacting with the cavity field is then given by \begin{equation} \label{Pem} P_{em}=|T_{b,n+1}|^2+|R_{b,n+1}|^2 \end{equation} All these probabilities are strongly dependent on the cavity mode profile. Their calculation needs to solve the one-dimensional time-independent Schr\"odinger equation. L\"{o}ffler \textit{et~al.}~\cite{Lof97} have calculated them as a function of the interaction length $\kappa_{n}L$ ($L$ is the cavity length and $\kappa_{n}=\kappa \sqrt[4]{n+1}$ with $\kappa=\sqrt{2mg/\hbar}$, and $m$ the atomic mass) for various cavity mode functions: mesa, $\textrm{sech}^2$ and sinusoidal modes. For the two first modes, analytical results of the probability $P_{em}(\kappa_{n}L)$ have been given. For the sinusoidal modes, detailed Wentzel-Kramers-Brillouin solutions have been presented. Nevertheless, Retamal \textit{et~al.}~\cite{Ret98} have shown that the WKB approximation may lead to inaccurate predictions for actual interaction and cavity parameters. The consideration of actual interaction and cavity parameters is a difficult numerical task as they can correspond to very large values of $\kappa_{n}L$. When no analytical solutions of the Schr\"odinger equation are available, usual numerical integration methods do not converge rapidly or do not converge at all. In this paper, we present a novel approach for solving the one-dimensional time-independent Schr\"odinger equation. Our method is described in Sec.~\ref{MethodDescription}. It may be applied for any potential function. It is efficient, stable and succeeds when other numerical integration methods fail. It may replace and improve advantageously the WKB approach. Details of the implementation of our method and validity tests are presented in Sec.~\ref{MethodImplementation}. We apply our method for calculating some new $P_{em}(\kappa_{n}L)$ curves. In particular, the Gaussian potential has been considered, thinking in open cavities in the microwave or optical field regime. These results are presented in Sec.~\ref{NewResults}. \section{Description of the method} \label{MethodDescription} The one-dimensional time-independent Schr\"odinger equation can be written in atomic units in the form \begin{equation} \label{SchrEq} \left( \frac{1}{2}\frac{d^2}{dx^2}+(E-V(x))\right)\phi(x)=0 \end{equation} We assume here that the potential $V(x)$ has non zero values only in a given region of the $x$ axis, say $\left[x_a,x_b\right]$. This region is divided into $J$ grid points, denoted $x_1,\ldots,x_J$ such that $x_a=x_1<x_2<\cdots<x_{J-1}<x_J=x_b$. Let us note $I_j$ ($0<j<J$) the region $\left[x_j,x_{j+1}\right]$, $I_0$ $\left]-\infty,x_1\right]$, and $I_J$ $\left[x_J,+\infty\right[$. The potential $V(x)$ is approximated in each region $I_j$ by a straight line connecting $V(x_j)$ and $V(x_{j+1})$. The approximated potential is noted $V_{approx}(x)$. Then Schr\"odinger equation takes the simple form in $I_j$ ($0 \leq j \leq J$): \begin{equation} \label{ApproxSchrEq} \left( \frac{d^2}{dx^2}+(a_j+b_{j}x)\right)\phi_{j}(x)=0 \end{equation} with $a_0=a_J=2E$, $b_0=b_J=0$, and for $0<j<J$ \begin{equation} \label{ajbjDef} \left\{ \begin{array}{l} a_j=2 \left(E-\frac{x_{j+1}V_{j}-x_{j}V_{j+1}}{x_{j+1}-x_{j}}\right)\\ b_j=-2 \frac{V_{j+1}-V_{j}}{x_{j+1}-x_{j}} \quad \textrm{with} \quad V_i=V(x_i),i=j,j+1 \end{array} \right. \end{equation} The most general solutions of Eq.~(\ref{ApproxSchrEq}) are given by \begin{equation} \label{phij} \phi_{j}(x)=C_{j}f_{j}^{+}(x)+D_{j}f_{j}^{-}(x) \end{equation} where $C_j$ and $D_j$ are two complex constants and $f_{j}^{+}(x)$ and $f_{j}^{-}(x)$ 2 functions depending on the $a_j$ and $b_j$ values. These functions are given in Table~\ref{FunctionsTable}. As shown in Table~\ref{FunctionsTable}, the functions $f_{j}^{+}(x)$ and $f_{j}^{-}(x)$ depend on the sign of $a_j+b_{j}x$. To avoid a change of this sign in a given region $I_j$, the set $x_1,\ldots,x_J$ must contain the roots of the equation $V(x)=E$. Thus, the condition $a_j+b_{j}x>0$ (resp. $a_j+b_{j}x<0$) is equivalent to that $E>V(x)$ (resp. $E<V(x)$) for $x \in I_j$. The complex constants $C_j$ and $D_j$ in Eq.~(\ref{phij}) are determined by use of the wavefunction asymptotic behaviour knowledge ($(C_0,D_0)$ or $(C_J,D_J)$ are supposed to be known) and of the set of conditions imposing the continuity of the wavefunction and its derivative along the $x$ axis~: \begin{equation} \label{ContinuityConditions} \left\{ \begin{array}{l} \phi_{j}(x_j)=\phi_{j-1}(x_j) \\ \phi'_{j}(x_j)=\phi'_{j-1}(x_j) \end{array} \right. \quad (0 < j \leq J). \end{equation} If $(C_0,D_0)$ are given, Eq.~(\ref{ContinuityConditions}) may be written as \begin{equation} \label{ContinuityConditionsA} \left( \begin{array}{l} C_{j} \\ D_{j} \end{array} \right) = A_{j}(x_j) \left( \begin{array}{l} C_{j-1} \\ D_{j-1} \end{array} \right) \quad \textrm{for} \: 0 < j \leq J, \end{equation} with \begin{eqnarray} A_j & = & \frac{1}{f_{j}^{+}g_{j}^{-}-f_{j}^{-}g_{j}^{+}} \times \nonumber\\ & & {}\left( \begin{array}{cc} f_{j-1}^{+}g_{j}^{-}-f_{j}^{-}g_{j-1}^{+} & f_{j-1}^{-}g_{j}^{-}-f_{j}^{-}g_{j-1}^{-} \\ f_{j}^{+}g_{j-1}^{+}-f_{j-1}^{+}g_{j}^{+} & f_{j}^{+}g_{j-1}^{-}-f_{j-1}^{-}g_{j}^{+} \end{array} \right) \end{eqnarray} and \begin{equation} g_i^{\pm}(x)=\frac{df_i^{\pm}}{dx} \, , \,i=j,j-1. \end{equation} If $(C_J,D_J)$ are known, Eq.~(\ref{ContinuityConditions}) should be written \begin{equation} \label{ContinuityConditionsB} \left( \begin{array}{l} C_{j-1} \\ D_{j-1} \end{array} \right) = B_{j}(x_j) \left( \begin{array}{l} C_{j} \\ D_{j} \end{array} \right) \quad \textrm{for} \: J \geq j > 0, \end{equation} with \begin{eqnarray} B_j & = & \frac{1}{f_{j-1}^{+}g_{j-1}^{-}-f_{j-1}^{-}g_{j-1}^{+}} \times \nonumber\\ & & {}\left( \begin{array}{cc} f_{j}^{+}g_{j-1}^{-}-f_{j-1}^{-}g_{j}^{+} & f_{j}^{-}g_{j-1}^{-}-f_{j-1}^{-}g_{j}^{-} \\ f_{j-1}^{+}g_{j}^{+}-f_{j}^{+}g_{j-1}^{+} & f_{j-1}^{+}g_{j}^{-}-f_{j}^{-}g_{j-1}^{+} \end{array} \right). \end{eqnarray} The knowledge of $\phi_{j}(x)$ in each region $I_j$ defines an approximated wavefunction of the Schr\"odinger equation (\ref{SchrEq}). The grid point number $J$ fixes the accuracy of the method. The higher $J$ is, the more accurate is the approximated potential $V_{approx}(x)$ and the better is the approximated wavefunction. The actual number $J$ to be considered for a given accuracy depends on various parameters (potential form, energy $E$, size of the region $\left[x_a,x_b\right]$). We show in Sec.~\ref{MethodImplementation} and \ref{NewResults} that a few hundreds are typical numbers. The wavefunctions yielded by our method are not divergent at the classical turning points (where $E = V(x)$), on the contrary of the WKB method. Although some Bessel functions of Table~\ref{FunctionsTable} are divergent at these points (as $z_j(x) \equiv a_j + b_j x = 0$), all the functions $f_{j}^{\pm}(x)$ and their derivative $g_{j}^{\pm}(x)$ admit finite limits there (see Table~\ref{Limits}). The transmission and reflection complex amplitudes (denoted respectively $t$ and $r$) of a particle incident upon a potential $V(x)$ may be calculated as well (and consequently the induced emission probability (\ref{Pem})). If we consider the outgoing wavefunction $\phi_{J}(x)=e^{ikx}$ with $k=\sqrt{2E}$, we can calculate the corresponding incoming wavefunction $\phi_{0}(x)=C_0\cos(kx)+D_0\sin(kx)$ by use of relations~(\ref{ContinuityConditionsB}). We then have \begin{equation} \label{tAndr} t=\frac{2}{C_0-iD_0} \quad , \quad r=\frac{C_0+iD_0}{C_0-iD_0} \end{equation} \section{Method implementation} \label{MethodImplementation} We have implemented the method described above on a PC Pentium based computer. Bessel functions of Table~\ref{FunctionsTable} have been coded using algorithms given by Zhang and Jin~\cite{Zha96}. The evaluation of the Mathematica $^{\circledR}$ Bessel functions has been discarded due to poor performances (thousand times slower than direct computation). For usual cavity parameters, $C_0$ and $D_0$ coefficients of Eq.~(\ref{tAndr}) can rapidly become very large. To avoid computation overflow errors, we have adopted a custom number representation. Our internal range of values was 9.9~E~$\pm$~($\sim$9~E~18) with 15 significants. As a first test of our method, we checked its ability to converge when the grid point number $J$ increases. We have calculated the induced emission probability $P_{em}$ of an atom interacting with a sech${}^2$ mode profile cavity ($u(x)=\textrm{sech}^2(x/L)$) as a function of the number $J$ and for a fixed interaction length $\kappa_{n}L$. Fig.~\ref{FigurePem_JSech2} presents the curve obtained in the case of $k/\kappa_n=0.01$ and $\kappa_nL=10$. This curve shows very clearly that $P_{em}$ converges to a given value as $J$ increases. Also, in that case, a $J$ number of 200 is sufficient for predictions to the accuracy of the graphics. For this calculation, the region $\left[x_a,x_b\right]$ was limited to 16 times the cavity length $L$ ($\sim 8$ times the FWHM of the potential function). In order to improve the convergence of our method with the number $J$, we renormalized the approximated potential so that the area under it remained constant whatever value $J$ had. Instead of using $V_i=V(x_i)$ in Eq.~(\ref{ajbjDef}), we have used $V_i=\alpha V(x_i)$ with $\alpha$ equal to the ratio of the area under $V(x)$ over the area under $V_{approx}(x)$. The study of the sech${}^2$ mode profile is a good test for a numerical method as there exists an analytical expression for $P_{em}(\kappa_nL)$. In this case we have (see L\"offler \textit{et~al.}~\cite{Lof97}) \begin{subeqnarray} \label{PemAnalSech2} t_n^{\pm} & = & \frac{\displaystyle \Gamma\left[1/2-i(kL+\xi_n^{\pm})\right] \: \cdot \: \Gamma\left[1/2-i(kL-\xi_n^{\pm})\right]}{\displaystyle \Gamma\left[-ikL\right] \: \cdot \: \Gamma\left[1-ikL\right]} \nonumber \\ & & \\ r_n^{\pm} & = & \frac{\displaystyle \Gamma\left[ikL\right] \: \cdot \: \Gamma\left[1-ikL\right]}{\displaystyle \Gamma\left[1/2+i\xi_n^{\pm}\right]\: \cdot \: \Gamma\left[1/2-i\xi_n^{\pm}\right]}t_n^{\pm} \end{subeqnarray} with $\xi_n^{\pm}=\sqrt{\pm(\kappa_nL)^2-1/4}\,$. Fig.~\ref{FigurePem_knLSech2} shows two $P_{em}(\kappa_nL)$ curves calculated for $k/\kappa_n=0.01$ and $k/\kappa_n=0.1$ in the case of the $\textrm{sech}^2$ mode profile. On this figure, the solid lines represent the analytical results deduced from formulas~(\ref{TRinExpr}), (\ref{Pem}) and (\ref{PemAnalSech2}) while the dotted ones represent those obtained using our method (the grid point number $J$ was fixed to 200). The agreement between these results is very good. \section{New results} \label{NewResults} \subsection{Fundamental sinusoidal mode} The fundamental sinusoidal mode profile \begin{equation} u(x)= \left\{ \begin{array}{ll}\sin(\pi x /L) \quad & \mathrm{for} \quad 0 < \mathnormal{x} < L \\ 0 & \mathrm{elsewhere} \end{array} \right. \end{equation} has been studied by L\"offler \textit{et~al.}~\cite{Lof97} and Retamal \textit{et~al.}~\cite{Ret98}. Their conclusions are not in agreement with regard to the behaviour of the induced emission probability for ultracold atoms ($k/ \kappa_n = 0.01$) interacting with large cavities ($\kappa_nL$ of the order of $10^5$). Such cavity parameters are of a great interest as they correspond to realistic values for Rydberg atoms interacting with microwave cavities in recent experiments performed by the Ecole Normale Sup\'erieure Group \cite{Bru96} (see discussion about the orders of magnitude in \cite{Ret98}). Retamal \textit{et~al.}~\cite{Ret98} have predicted well resolved resonances in the curve $P_{em}(\kappa_nL)$ for the parameters cited above, whereas the curve predicted on the basis of the results of L\"offler \textit{et~al.}~\cite{Lof97} looks very different. We have calculated the $P_{em}(\kappa_nL)$ curve for an atom interacting with this mode profile by use of our method. Fig.~\ref{FigurePem_knLFundSin} presents this curve for $\kappa_nL$ comprised between 100000 and 100010 ($k/ \kappa_n = 0.01$ and $J=100$). Our curve shows the same well resolved resonances as those predicted by Retamal \textit{et~al.} \cite{Ret98}. So we confirm their predictions that the WKB approximation cannot be used when considering ultracold atoms with $k/\kappa_n = 0.01$ and realistic interaction lengths of the order of $10^5$. Our calculations confirm also that these resonances have got smeared at $\kappa_nL=10^5$ if we consider warmer atoms characterized by $k/ \kappa_n = 0.1$. For this last value of $k/ \kappa_n$ the WKB approximation still holds and $P_{em}$ is then invariably equal to $1/2$. It is good to remark here the importance of still having resolved resonances in $P_{em}$ for actual interaction and cavity parameters, as this result will permit the possibility of testing, in principle, the mazer effect with present cavities. The construction of a re-entrant cavity, as it was proposed in Ref. \cite{Lof97}, would not be necessary for testing the mazer effect, but for improving the resolution of the resonances by approximating a mesa function mode profile. \subsection{First excited sinusoidal mode} The first excited sinusoidal mode profile \begin{equation} u(x)= \left\{ \begin{array}{ll}\sin(2\pi x /L) \quad & \mathrm{for} \quad 0 < \mathnormal{x} < L \\ 0 & \mathrm{elsewhere} \end{array} \right. \end{equation} has been studied by L\"offler \textit{et~al.}~\cite{Lof97}. These authors have derived an expression for $P_{em}$ on the basis of a detailed WKB calculation. They have shown that, for large interaction lengths (without being very explicit about the word ``large''), the behaviour of the induced emission probability is described by $P_{em}(\kappa_nL)=\sin^2(\Delta_n)$ with \begin{equation} \label{Deltan} \Delta_n = \kappa_nL \cdot \frac{1}{\pi}\int_0^{\pi/2}\!\!\!\!\sqrt{(k/\kappa_n)^2 + \cos(x)}dx \end{equation} which can be written $\Delta_n = \kappa_nL \cdot Const.$ if $k/\kappa_n$ is a fixed ratio. In Fig.~\ref{FigurePem_knLFirstSin}, we present two curves of $P_{em}(\kappa_nL)$ calculated for $\kappa_nL$ comprised between 100000 and 100020 (for $k/\kappa_n = 0.01$ and $k/\kappa_n = 0.1$). The first curve does not exhibit a square sine dependence over the interaction length, whereas the second one does. This indicates that $\kappa_nL=10^5$ cannot be considered as ``large'' in the sense of L\"offler {\it {et al.}}~\cite{Lof97} for $k/\kappa_n = 0.01$. For $k/\kappa_n = 0.1$, $\Delta_n = 2\pi \kappa_nL / T$, with the period $T$ equals to $\sim\!\!16.3$ (according to Eq.~(\ref{Deltan})). This period is well reproduced by our calculations (see curve (b) on Fig.~\ref{FigurePem_knLFirstSin}). It is a remarkable result of L\"offler {\it {et al.}} \cite{Lof97} the possibility of building state-changing and state-preserving mirrors for atoms by modifying the length of the cavity using this first excited sinusoidal mode. From Fig.~\ref{FigurePem_knLFirstSin} we conclude that resonances in $P_{em}$ for the first excited mode are even narrower than those predicted by a square sine function $\sin^2(\Delta_n)$ . This is a very convenient result as we are considering actual interaction and cavity parameters. \subsection{Gaussian mode} Up to-date, the Gaussian cavity mode profile has not been studied exactly in the quantum theory of the mazer. L\"offler \textit{et~al.}~\cite{Lof97} have argued that the $\textrm{sech}^2$ mode profile could be used as a good approximation of the Gaussian one. To verify this assumption, we have calculated various $P_{em}(\kappa_nL)$ curves for the profile \begin{equation} \label{gaussianProfile} u(x)=e^{-\frac{x^2}{2 \sigma^2}} \end{equation} The parameter $\sigma$ was fixed to $\sqrt{2/ \pi} L$ in order to adopt the same normalization factor for the two profiles (identical area under the modes). Fig.~{\ref{FigurePem_knLGaussian}} shows our results for $k/\kappa_n=0.1$ in the range $\kappa_nL=0$ to $\kappa_nL=20$. Qualitatively both profiles exhibit the same behaviour~: the resonances in the curves get smeared with increasing values of $\kappa_nL$. But this phenomenon is not so marked in the case of the Gaussian profile. Resonances still exist for longer interaction lengths. This is not a surprising result as the Gaussian profile is growing more abruptly than the $\textrm{sech}^2$ one. Thus it is in some sense ``closer'' to the mesa mode, which exhibits resonances at infinity. We have also considered the case $k/\kappa_n=0.01$. Our calculations have shown that 90\% damped oscillations are still present in the $P_{em}(\kappa_nL)$ curve for interaction lengths approximately 3 times larger in comparison with the $\textrm{sech}^2$ mode case. For these calculations, the region $[x_a,x_b]$ was limited to 16 times the cavity length $L$ and the grid point number $J$ was fixed to 300. As it was pointed out in Ref. \cite{Lof97} the mazer effect is not restricted to the microwave domain and it might be tested more efficiently in the optical domain. Note that typical optical cavities have a Gaussian mode profile and it is possible to consider large coupling constants, fact that will help for testing the mazer effect. We want to call attention to the fact that Hood {\it et al.}, in Ref. \cite{Hoo98}, presented the first experimental result for which the interaction energy $\hbar g$ is greater than the atomic kinetic energy ($k/\kappa_n \ll 1$). \section{Summary} We have developed a new method for solving one-dimensional scattering problems that may be applied advantageously instead of the WKB approach. This has enabled us to calculate efficiently the induced emission probability $P_{em}$ of an atom interacting with a high-$Q$ cavity for various mode profiles. Two sinusoidal modes have been considered. For these cases, we have been able to assert that the WKB approximation cannot be used in the computation of $P_{em}$ for ultracold atoms ($k/ \kappa_n = 0.01$) interacting with actual cavities characterized by $\kappa_nL$ of the order of $10^5$. Significant and convenient different physical predictions are found, if we compare our results with previous works \cite{Lof97}. The Gaussian mode profile has also been considered and we have shown that, although it exhibits a similar behaviour in comparison with the $\textrm{sech}^2$ mode profile, the resonances in the $P_{em}(\kappa_nL)$ curves exist for significantly larger values of $\kappa_nL$. The Gaussian mode is relevant when considering open cavities in the microwave or optical domains. The presented numerical method will be helpful for computing the induced emission probability in the case of the recently studied two-photon mazer \cite{Zha99}, when considering field mode profiles different from the mesa function. This and other applications of this numerical method will be presented elsewhere. \acknowledgements This work has been supported by the Belgian Institut Interuniversitaire des Sciences Nucl\'eaires (IISN) and the Brazilian Conselho Nacional de Desenvolvimento Cient\'{\i}fico (CNPq). E. S. wants to thank Prof. Nicim Zagury for helpful comments and suggestions.
\section{Introduction} The mass of a galaxy is a fundamental quantity in understanding its dynamics and structure. Mass distribution in galaxies has been extensively studied with optical and HI rotation curves. Several studies revealed that spirals galaxies have flat rotation curves even at the observed outermost points, indicating the existence of extended dark halos (e.g., Sancisi \& van Albada 1987) . The extent and total mass of dark halos are, however, not understood well and yet to be studied in detail. For further investigation of extended halos, different approaches are required to trace the mass distribution beyond the HI disk, where rotation curves cannot be measured. Binary galaxies are useful for the determination of the total mass or mass-to-light ratio ($M/L$) of galaxies, like stellar masses are measured from the motion of binary stars. Unlike stellar binaries, however, the total mass of individual binary cannot be directly determined because of the long orbital periods. Instead, statistical treatment is necessary to obtain the average mass or $M/L$ of the sample galaxies. Many efforts were made to determine the $M/L$ ratio of binary galaxies statistically (e.g., Page 1952; Karachentsev 1974; Turner 1976a, b; Peterson 1979a, b; White 1981; van Moorsel 1987; Schweizer 1987; Chengalur et al.1993; Soares 1996). In binary galaxy studies, a careful selection of binary galaxies is very important, since the biases in selecting pairs should be corrected for to determine the mass or $M/L$ ratio. Turner (1976a) proposed well-defined selection criteria based only on the positions and magnitudes of galaxies. Later investigations (e.g., Peterson 1979a) also made use of similar selection criteria independent of radial velocities, which are so-called {`}velocity-blind{'} pairs. According to such velocity-blind selection criteria, two galaxies are regarded as a pair if they have no close companion compared to their projected separation. This {`}velocity-blind{'} selection criterion is simple and convenient for pair selection, but its problem is that the criterion could introduce strong bias toward pairs with small separations; for pairs with wider separations, company galaxies are searched for in a larger region, leading to exclusion of widely-separated pairs with higher probability. In fact, the average separations of selected pairs in these studies were 50 kpc $\sim$ 100 kpc (see Peterson 1979b). Since the dark halos could extend beyond this range, it is important to study binary galaxies further based on pairs with wider separations. The other major problem of the velocity-blind sample is that the sample suffers from the contamination of {`}optical pairs{'}, which consists of two isolated galaxies projected close by chance. In order to reduce the contamination of optical pairs, it is better to select binary galaxies based not only on the positions but also on the radial velocities. Fortunately, the number of radial velocity observations is rapidly increasing thanks to recent large-scale redshift surveys. Moreover, the observational uncertainty has been significantly reduced due to the recent development of observational instruments, which enables us to estimate $M/L$ with better accuracy compared to previous studies. Therefore, it is interesting to study binary galaxies again by utilizing such huge data. For these reasons, in this paper we study the mass-to-light ratio of binary galaxies by making use of the database. The plan of this paper is as follows. In section 2, we will describe how to select widely-separated pairs effectively, while reducing the contamination of optical pairs. The selection criteria, and the basic data for selected pairs will be presented in section 2. In section 3 we will perform maximum-likelihood analysis based on the orbital models of binary galaxies, and determine $M/L$. We will also consider $M/L$ dependence on galaxies{'} type. Discussions on the dark halo extent will be given in section 4. \section{Selection of Pairs} \subsection{Basic Idea, Selection Criteria, and Sample} The observable quantities for orbital motion of pairs are the projected separation $r_{\rm p}$, and the radial velocity differences $v_{\rm p}$. A set of $r_{\rm p}$ and $v_{\rm p}$ can be used to estimate the total mass of a pair through an estimator of mass, for example, $r_{\rm p} v_{\rm p}^2/G$. However, the mass of galaxies varies by about 3 order of magnitude from dwarf galaxies to giant ellipticals. A better quantity which represents the mass of galaxies is the mass-to-light ratio, $M/L$. The mass-to-light ratio of galaxies is expected to vary much less than the mass itself, and hence we focus on the mass-to-light ration of pairs in this paper. What can be obtained through binary galaxy analysis is the total mass to total light ratio of pairs which is written as $(M_1+M_2)/(L_1+L_2)$, but in the rest of this paper we denote this ratio as $M/L$ for simplicity. Note that if $M_1/L_1= M_2/L_2$, the total mass-to-light ratio $M/L$ is equal to $M_1/L_1$ and $M_2/L_2$. For convenience in $M/L$ estimate, we define the luminosity-corrected separation $R_{\rm p}$ and the luminosity-corrected velocity difference $V_{\rm p}$ as \begin{equation} R_{\rm p} \equiv r_{\rm p}/L^{1/3}, \end{equation} \begin{equation} V_{\rm p} \equiv |v_{\rm p}|/L^{1/3}. \end{equation} A combination of $R_{\rm p}$ and $V_{\rm p}$ can give an estimator of the mass-to-light ratio of pairs. This estimator, which we call the projected mass-to-light ratio, is defined as \begin{equation} (M/L)_{\rm p} \equiv \frac {r_{\rm p} v_{\rm p}^2}{G L} = \frac{R_{\rm p} V_{\rm p}^2}{G}. \end{equation} If a bound pair of galaxies are separated so widely that they can be approximated as two point masses, the law of energy conservation gives that \begin{equation} \frac{r v^2}{2G L}\le M/L, \end{equation} because the total energy for a bound pair is always negative. A combination of equation (3) with inequality (4) gives \begin{equation} (M/L)_{\rm p} \le 2 (M/L). \end{equation} If all pairs are bound and have the same $M/L$, the binary populations in the $R_{\rm p}$-$V_{\rm p}$ phase space lie below the envelope which corresponds to $2(M/L)$. In practice, the number of pairs is limited and insufficient to see the true envelope corresponding to $2 (M/L)$ in the $R_{\rm p}$-$V_{\rm p}$ space. Detailed calculations of probability distribution show that pairs are likely to concentrate to small $V_{\rm p}$, and thus smaller $(M/L)_{\rm p}$ due to projection effect(e.g., Noerdlinger 1975). We will discuss this in later section by calculating the probability distribution based on the Monte-Carlo simulation. In any case, the pair distribution in the $R_{\rm p}$-$V_{\rm p}$ phase space can be used for testing whether or not bound pairs are efficiently selected: while bound pairs are likely to have small $V_{\rm p}$, optical pairs could have extremely high $V_{\rm p}$ and $(M/L)_{\rm p}$. Here we describe the selection criteria for pairs. For convenience, we define the total luminosity of a pair normalized with $10^{10}L_\odot$ in the B band as $L_{10}\equiv (L_1+L_2)/(10^{10}L_\odot)$. Note that this luminosity roughly corresponds to that of the Milky Way Galaxy. We re-define $R_{\rm p}$ and $V_{\rm p}$ normalizing with luminosity $L_{10}$ as, \begin{equation} R_{\rm p} = \frac{r_{\rm p}}{L_{10}^{1/3}} \;{\rm (kpc)}, \end{equation} and \begin{equation} V_{\rm p} = \frac{|v_{\rm p}|}{L_{10}^{1/3}} \;{\rm (km\; s^{-1})}. \end{equation} As the first step of pair selection, we have to select a pair of galaxies that are relatively close to each other both in the sky plane and in the redshift space so that they are likely to be bound. Since the average separations in previous studies were around 100 kpc, and since we are interested in widely-separated pairs, we set the maximum projected separation of a pair in the sky plane as $$ 1: R_{\rm p} \le 400 \;{\rm kpc}. $$ The maximum velocity difference must also be large enough to include pairs orbiting around each other at high velocity. Since galaxies as luminous as $10^{10}L_\odot$ have rotation velocity of $\sim$ 200 km s$^{-1}$, the velocity difference of a pair could be as high as a few hundred km s$^{-1}$. Hence, we set the maximum velocity difference of pairs as $$ 2: V_{\rm p} \le 400 \;{\rm km\; s^{-1}}. $$ Note that the radial velocity difference is usually quite small compared to the true velocity difference due to projection effect (Noerdlinger 1975). Thus, physical pairs are unlikely to have larger radial velocity differences than the maximum value given above. The observational data set cannot be complete to faint galaxies, and hence we limit the application of our analysis to sufficiently bright galaxies. Since nearby galaxies are cataloged almost completely down to 15.5 magnitude (e.g, de Vaucouleurs et al. 1991), we set a criterion for the B band magnitude as $$ 3: m_1, m_2 \le 15.0 \;{\rm mag}, $$ and also set an upper limit for the total magnitude of a pair in the B band, $m_{1+2}$, as $$ 4: m_{1+2} \le 13.5 \;{\rm mag}$$ In order for selected pairs to be likely to be bound, pair galaxies must be isolated well. We regarded two galaxies as a pair if all of its companion galaxies brighter than $m_{1+2} + 2.0 \;{\rm mag}$ satisfy both $$ 5: \frac{r_i}{L_{10}^{1/3}} \ge a\; 400 \; {\rm kpc},$$ and $$ 6: \frac{|v_i|}{L_{10}^{1/3}} \ge b\; 400 \; {\rm km s^{-1}}.$$ Here $r_i$,and $v_i$ are the projected separation and the radial velocity difference of $i$th companion galaxy with respect to the luminosity center of the pair. Note that we set the lower limit of the total magnitude $m_{1+2}$ to be 13.5 mag (criterion 4). The faintest galaxies that should be considered is, hence, at 15.5 mag, to which magnitude galaxies are cataloged almost completely. Parameters $a$ and $b$ determine the volume for companion search, and so determine the degree of isolation. Note that the volume depends only on the total luminosity of a pair, $L_{10}$, but independent of the separation $R_{\rm p}$ of a pair. Therefore, as far as pairs with the same luminosity are concerned, companions are searched for in the same volume, and thus the criterion does not introduce bias toward pairs with small separations. We set $b=1.5$ throughout this paper, but tested three values of $a$ (1.5, 2.0, and 2.5) to seek a value for effective selection of bound pairs. We applied criteria described above to the sample of galaxies that we compiled for this study using NED (NASA Extra-galactic Database). The sample consists of bright nearby galaxies with redshift less than 4,500 $\rm km \; s^{-1}\;$. The upper limit for redshift is introduced because the number of bight pairs which satisfy the criteria 3 and 4 becomes small at large redshift. The data for positions, heliocentric velocities, and the B band magnitudes were mainly taken from NED, and supplied with RC3 catalog (de Vaucouleurs et al.1991). The distances to the pairs are obtained using the redshift of the luminosity center ($H_0=$ 50 $\rm km \; s^{-1}\;$ is assumed). In order to avoid the error in distance due to the local deviation from Hubble flow, galaxies with redshift smaller than 1,000 $\rm km \; s^{-1}\;$ are excluded from the sample. Galaxies in clusters and close to clusters may deviate from the Hubble flow even beyond the redshift of 1000 $\rm km \; s^{-1}\;$, but this effect is expected to be small because the pairs selected with criteria 5 and 6 are likely to be field binary galaxies. Galaxies with $b \le 20^\circ$ are also excluded from the sample since objects at low galactic latitude may be significantly obscured by galactic extinction The sample we compiled consists of 6475 galaxies with magnitude brighter than 15.5 mag and redshift between 1000 km s$^{-1}$ and 4500 km s$^{-1}$. The uncertainty in the magnitude is typically 0.2 mag., which leads to the uncertainty in $L$ of 20 \%. The corrections for intrinsic absorption and galactic extinction were made according to de Vaucouleurs et al.(1991). The sample is, of course, incomplete in terms of redshift because redshift measurements were not made for all galaxies. This incompleteness leads to possible mis-identification of pairs, if the criteria described above are applied only to a sample of redshift-know galaxies. To correct for the effect of redshift incompleteness, primary binary candidates are at first searched in the sample of redshift-known galaxies, and then a redshift-blind search was performed for the primary binary candidates. If there is any redshift-unknown companion which is brighter than $m_{1+2} + 2.0$ mag and is so close to the pair that the criterion 5 is violated, the pair was rejected from the binary candidates. About 30 \% of pairs in the primary binary candidates were rejected through this procedure. The sample of binary galaxies after the correction for the redshift incompleteness still contains some pairs that are not appropriate for this study. For instance, the basic data for the analysis, such as $v_{\rm p}$, $m_1$, and $m_2$ could be quite uncertain for some pairs. In particular, the uncertainty in the radial velocity is crucial for $M/L$ determination, as the $M/L$ estimator depends on $V_{\rm p}^2$. Therefore, if redshift uncertainty is not reported for any of the two galaxies of pair, the pair is excluded. This process reduced the number of pairs by 7\%. If the magnitude uncertainty and the absorption-corrected magnitude are not available, the pair is also excluded, and in this process 7\% of primary binary candidates were rejected. Moreover, a galaxy could appear in the binary sample twice or more with different partner. This can happen if one of pair galaxies is a bright galaxy like the cD galaxy and it has several companion galaxies around it. In this case, however, these galaxies should be regarded as cluster or group rather than binary galaxies. Therefore, we also excluded possible clusters or groups of galaxies that appear in the binary sample twice or more. We found only two possible groups in the primary binary candidates. \subsection{Results} Figures 1 show the distribution of thus selected binary galaxies in the $R_{\rm p}$-$V_{\rm p}$ phase space. Two cases for the isolation parameter, $a$=1.5 and 2.5, are shown. The number of selected pairs is 109 and 57, respectively. In the case of $a$=1.5, the pairs in the $R_{\rm p}$-$V_{\rm p}$ space shows only weak concentration toward small $V_{\rm p}$, and a large number of galaxies have high $(M/L)_{\rm p}$ exceeding a few hundred $M_\odot/L_\odot$. Even if their true $M/L$ is a few hundred, it is unlikely that so many galaxies appear to have so large $(M/L)_{\rm p}$ in the projected phase space, as $(M/L)_{\rm p}$ is expected to be significantly smaller than the true $M/L$ due to projection effect (Noerdlinger 1975; see also Section 3 of the present paper). This indicates that they are probably optical pairs, and that the degree of isolation is not strong enough to select bound pairs effectively. On the other hand, for $a=2.5$, the concentration of pairs to $V_{\rm p}$=0 is much more clear than for $a=1.5$. Most of 57 pairs are distributed below $(M/L)_{\rm p}$ of 20 in solar unit, and there are only few galaxies that have high $(M/L)_{\rm p}$. This correlation between $R_{\rm p}$ and $V_{\rm p}$ are naturally explained if the separation of pairs are larger than the extent of halos so that the pairs can be approximated as point masses (but note that even in case of extended hale such an envelope would appear in the projected phase-space like point-mass cases; see Soares 1990). In the rest of this paper, we use the binary galaxies sample selected with $a=1.5$ and 2.5 for $M/L$ determination. We call the sample selected with $a=2.5$ as sample I, and the one selected with $a=1.5$ as sample II. Table 1 summarizes the basic data for 57 pairs in sample I. \section{$M/L$ determination} In this section, we estimate the $M/L$ ratio of the sample pairs selected above. We construct the orbital models for physical pairs, and calculate the probability distribution of pairs in the $R_{\rm p}$-$V_{\rm p}$ phase space considering the contamination of optical pairs. Then, we compare the models with the observational data, and determine the $M/L$ ratio based-on maximum-likelihood analysis. \subsection{Distribution of Bound Pairs} First we construct models for orbital populations of binaries. For simplicity, binary galaxies are treated as point masses in the following analysis. As can be seen in figure 1b, pairs show strong concentration toward small $V_{\rm p}$ in the $R_{\rm p}$-$V_{\rm p}$ space, which is just expected from the point-mass assumption. Further tests for validity of the assumption will be made in the next section. An ensemble of well-mixed binary population satisfy the Jeans equation (Binney and Tremaine 1987), \begin{equation} \frac{d (\nu \overline{v_{\rm r}^2})}{d r} + \frac{2\nu \beta \overline{v_{\rm r}^2}}{r}=-\frac{GM}{r^2}\nu. \end{equation} Here $\nu$ denotes the separation distribution of pairs, and $\beta$ is the anisotropy parameter defined as \begin{equation} \beta = 1 - \frac{\overline{v_\theta^2}}{\overline{v_{\rm r}^2}}, \end{equation} where $\overline{v_\theta}$ and $\overline{v_{\rm r}}$ denote each component of velocity ellipsoids. Note that $\beta = -\infty$ for circular orbits, $\beta = 0$ for isotropic orbits, and $\beta = 1$ for radial orbits. We may rewrite equation (8) by normalizing with luminosity as \begin{equation} \frac{d (\nu \overline{V_r^2})}{d R} + \frac{2\nu \beta \overline{V_r^2}}{R}=-\frac{G M}{R^2 L}\nu, \end{equation} where $R=r/L^{1/3}$ and $\overline{V_r^2}=\overline{v_r^2}/L^{2/3}$. Note that they are true separation and velocity, but not projected ones. For the separation distribution $\nu$, we also assume a power law with an inner cutoff radius $r_{\rm min}$ as, \begin{equation} \nu(R) \propto R^{-\gamma} \;\;\;\;\; {\rm for} \;\;\;\; R \ge R_{\rm min}. \end{equation} We introduced the cutoff radius because galaxies have finite sizes, and pairs that are too close are not likely to exist. The model used here is, therefore, not exactly a scale-free model (cf. White 1981). In order to model the distribution of pairs, one should choose suitable values for parameters $\beta$, $\gamma$, and $R_{\rm min}$. The parameters related to the separation distribution are obtained directly from observed separation distribution, because the probability distribution for projection effect can be written analytically as \begin{equation} p [R_{\rm p} | R] = \frac {2 R_{\rm p}}{\pi R(R^2 - R_{\rm p}^2)^{1/2}} \;\;\;\;\;\;\;\; ({\rm for\;\;\;} R_{\rm p} \le R). \end{equation} We compared the separation distribution of observed and model pairs, and obtained the best-fit values $\gamma = 2.6$ and $R_{\rm min}= 10$ kpc. In the rest of this paper, we adopt these best-fit values for $\gamma$ and $R_{\rm min}$, but we note the results are not sensitive to changes in the assumed values. Once the separation distribution is obtained, the distribution of the velocity difference is obtained by solving the Jeans equation [eq.(8)]. Then, one can calculate the probability distribution of pairs in the $R_{\rm p}$-$V_{\rm p}$ phase space, $p_{\rm bin} [R_{\rm p},V_{\rm p}|(M/L)]$, by taking the projection effect into consideration. \subsection{Selection Bias and Contamination of Optical Pair} In addition to the orbital models for true pairs, here we consider the selection effect for pairs and the contamination of optical pairs. As described in the previous section, the isolation criteria are independent of the separation or velocity difference of pairs, and hence, the sample is free from biases both for the separation in the sky plane, and for the separation along the line of sight. The isolation criteria, however, may cause possible exclusion of true pairs due to chance projection of another companion galaxy that are not physically related to the pair. According to the isolation criteria, the maximum velocity difference of a pair is $\sim 400$ $\rm km \; s^{-1}\;$ for galaxies with $L\sim 10^{10} L_\odot$. Therefore, a system of a true pair plus any foreground or background galaxy at distant within 8 Mpc from the pair cannot be regarded as a true pair because the criterion 6 is violated ($H_0 =50$ km s$^{-1}$ Mpc$^{-1}$ assumed). Unfortunately, there is no way to determine whether the observed velocity difference of two galaxies is due to Hubble flow or due to binary orbital motion, and so this kind of exclusion of true pairs is unavoidable. Furthermore, two galaxies which are separated well along the line of sight and are not physically associated could be regarded as a pair because of misidentification of the redshift as binary orbital motion. For these reasons, the sample selected in the present paper is far from perfect but likely to contain unphysical pairs which would lead to wrong estimates of $M/L$. Therefore, the exclusion of true pairs and the contamination of optical pairs must be taken into consideration for $M/L$ determination. Fortunately, the possibility of true-pair exclusion is independent of $R_{\rm p}$ or $V_{\rm p}$, as the selection criteria do not depend on them. Hence, the probability distribution of pairs in $R_{\rm p}$-$V_{\rm p}$ space, which is to be compared with the observed pairs, can be expressed as \begin{equation} p[R_{\rm p}, V_{\rm p}|M/L,f] = f p_{\rm bin}[R_{\rm p}, V_{\rm p}|(M/L)] + (1-f) p_{\rm opt}[R_{\rm p}, V_{\rm p}], \end{equation} where $f$ is a constant corresponding to the fraction of true pairs out of observed pairs. Clearly the first term on the right side expresses the contribution of true binaries, and the second term describes the contamination from optical pairs. Probabilities $p$, $p_{\rm bin}$, and $p_{\rm opt}$ are normalized so that \begin{equation} \int \int p\; dR_{\rm p}dV_{\rm p} = \int \int p_{\rm bin}\; dR_{\rm p}dV_{\rm p} = \int \int p_{\rm opt}\; dR_{\rm p}dV_{\rm p} = 1, \end{equation} where the integrations are performed from 0 to 400 kpc for $R_{\rm p}$, and from 0 to 400 $\rm km \; s^{-1}\;$ for $V_{\rm p}$. The possibility of mis-identification of optical pairs is proportional to the number density of galaxies. It is generally known that the distribution of galaxies in the Universe is not uniform but shows strong clustering, which is usually described in terms of the two-point correlation function (e.g. Peebles 1993). With this function the probability distribution of optical pairs in the $R_{\rm p}$-$V_{\rm p}$ phase space can be written as \begin{equation} p_{\rm opt}[R_{\rm p},V_{\rm p}] \;dR_{\rm p} dV_{\rm p}\propto \left[1+\xi(r)\right] R_{\rm p} \;dR_{\rm p} dV_{\rm p}, \end{equation} where $\xi(r)$ is the two-point correlation function, and this is usually written in the form of \begin{equation} \xi(r) = \left( \frac{r_0}{r} \right)^{q}. \end{equation} The two-point correlation function is determined well in the scale of 10 Mpc but less certain in the scale of 1 Mpc. Hence in the following analysis we consider two cases, no clustering case with $q=0$ and clustering case with $q=1.8$ and $r_0=10$ Mpc (Peebles 1993), and see how the $M/L$ estimates depend on the clustering effect. The separation $r$ can be calculated from the projected separation and the velocity difference by assuming the Hubble constant of 50 km s$^{-1}$ Mpc$^{-1}$. Note that in any case the probability distribution of optical pairs in the $R_{\rm p}$-$V_{\rm p}$ space is independent of the $M/L$ ratio of galaxies. \subsection{$M/L$ Determination} To evaluate the mass-to-light ratio and the fraction of true pairs we make use of the maximum-likelihood method for $M/L$ and $f$. The probability for finding a pair at $R_{\rm p}$ and $V_{\rm p}$ in the projected phase space is proportional to $p[R_{\rm p},V_{\rm p}|M/L,f]$, and hence the logarithmic likelihood for finding all observed pairs at their observed positions in the projected phase space is expressed as the summation of the probability for finding each pair at its position. Therefore, the logarithmic likelihood of $M/L$ for observed pairs can be written as \begin{equation} \log {\cal L} (M/L,f) = \sum n(R_{\rm p}, V_{\rm p}) \; \log p\,[R_{\rm p}, V_{\rm p}|M/L,f], \end{equation} where $n(R_{\rm p}, V_{\rm p})$ denotes the observed number of pairs having $R_{\rm p}$ and $V_{\rm p}$, and the summation is done over the whole projected phase space ($R_{\rm p}$ less than 400 kpc and $V_{\rm p}$ less than 400 km/s). Evidently $n(R_{\rm p}, V_{\rm p})$ is integral as long as the values of $R_{\rm p}$ and $V_{\rm p}$ are determined with sufficient accuracy. However, for the pairs we consider here $R_{\rm p}$ and $V_{\rm p}$ have uncertainties, and the uncertainty in $V_{\rm p}$ is particularly crucial for $M/L$ determination because the $M/L$ estimator depends on $V_{\rm p}^2$. Therefore, we treated each observed pair as a Gaussian distribution spread in the direction of $V_{\rm p}$, and then $n(R_{\rm p}, V_{\rm p})$ is given as \begin{equation} n(R_{\rm p}, V_{\rm p})\; dR_{\rm p} dV_{\rm p} \propto \sum_i g_i \; dR_{\rm p} dV_{\rm p}, \end{equation} where \begin{equation} g_i = (2\pi \sigma_i)^{-1/2} \exp \left(\frac{-(V_{\rm p}-V_i)^2}{2\sigma_i^2}\right). \end{equation} Here $V_i$ and $\sigma_i$ denote the observed $V_{\rm p}$, and the uncertainty for $V_{\rm p}$ for $i$th pair, respectively. Note that $n$ is normalized so that $\int \int n \; dR_{\rm p} dV_{\rm p}= N_{\rm tot}$, where $N_{\rm tot}$ is the total number of pairs. Since the probability distribution of physical pair in the $R_{\rm p}$-$V_{\rm p}$ phase space, $p_{\rm bin}[R_{\rm p}, V_{\rm p} | (M/L)]$, cannot be expressed analytically due to the projection effect, we performed Monte-Carlo simulations to evaluate $p_{\rm bin}$. The distribution of one million pairs in the projected phase space were simulated assuming random orientation of orbital planes with respect to the line of sight, and then the logarithmic likelihood (equation [16]) was calculated in the parameter space of $f$ and $M/L$. Figure 2 shows the likelihood contours for sample I (57 pairs) for the case of $q=0$ (no clustering for optical pairs). The thick lines are for $\beta =0$ (isotropic orbit) and dotted lines are for $\beta = -\infty$ (circular orbit). The figures show that the M/L estimates are not affected strongly by the assumed orbital parameters. The best estimates of $M/L$ are $35_{-5}^{+7}$ for $\beta = 0$ and $28_{-3}^{+5}$ for $\beta = -\infty$, with true binary fraction $f$ of $0.88_{-0.1}^{+0.07}$ for both cases (the error bars denote the 68 \% confidence level). The results for the true binary fraction $f$ indicates that most of pairs in sample I are likely to be bound. The expected number of optical pair is about 7 out of 57 pairs, which is comparable to the number of pairs which appear in Figure 1b above the envelope corresponding to $(M/L)_{\rm p}$ of 20. On the other hand, figure 3 shows the likelihood contours for sample I like figure 2, but for $q=1.8$ (clustering for optical pairs). The contours for two orbital models are shown, and again one can see that the weak dependence of $M/L$ on the orbital parameter. However, the true pair fractions $f$ are quite different from those for no clustering cases, as we obtained $f=0.71_{-0.15}^{+0.14}$ ($\beta = 0$) and $f=0.73_{-0.15}^{+0.14}$ ($\beta = -\infty$) for clustering case. This is because the expected number of optical pairs with small $V_{\rm p}$ is much larger than that for no clustering case, and hence more number of galaxies with small $V_{\rm p}$ are regarded as optical pairs. However, the best estimates of $M/L$ are $36_{-4}^{+8}$ for $\beta = 0$ and $30_{-3}^{+6}$ for $\beta = -\infty$, which are fairly close to those for no clustering cases. In figure 2 and 3 the best estimates of $M/L$ change little depending on $f$. This is because the $M/L$ is essentially determined by the pairs which lie below the envelope in figure 1b: in the most range of $f$ (e.g. $f$ less than 0.95) the pairs far above the envelope are likely to be optical pairs, and have little effect on the $M/L$ determination. On the other hand, if $f$ is set to be almost unity, the most likely $M/L$ could become as high as 100 to explain the pairs with high $(M/L)_{\rm p}$ without optical pairs. However, the likelihood for finding $f$ of almost unity is very small compared to that for $f$ between 0.6 to 0.9, for in that case the concentration of pairs to small $(M/L)_{\rm p}$ seen figure 1 cannot be explained at all. In order to test whether $M/L$ and $f$ obtained above can reproduce the distribution of observed pairs, we simulated distribution of model pairs in the $R_{\rm p}$-$V_{\rm p}$ phase space using the best-fit parameters. Figure 4 shows the simulated distribution of 57 modeled pairs with $f=0.88$ and $M/L=35$ assuming isotropic orbits for true pairs and no clustering for optical pairs. The figure resembles well the observed distribution in figure 1b in many aspects; small fraction of pairs with extremely high $(M/L)_{\rm p}$, concentration of pairs to small $V_{\rm p}$, and the envelope corresponding to $(M/L)_{\rm p}\sim 20$. This simulation confirms the validity of the results. In order to see if these results depend strongly on the sample we used, we also performed the same analysis for sample II (109 pairs), which are selected with weaker isolation criterion ($a=1.5$). As seen in Figure 1a, this sample is likely to contain more number of optical pairs than sample I due to the weak selection criterion. In fact, the resultant value of $f$ is 0.80$_{-0.08}^{+0.07}$ for no clustering case ($q=0$) and $0.62_{-0.11}^{+0.10}$ for clustering case ($q=1.8$), with $\beta=0$ for both cases. However, the best estimates $M/L$ are $35_{-3}^{+5}$ ($\beta=0$) and $28_{-2}^{+5}$ ($\beta=-\infty$) for no clustering cases, and $36_{-3}^{+12}$ ($\beta=0$) and $30_{-3}^{+8}$ ($\beta=-\infty$) for clustering cases. These results remarkably agree with those for sample I, indicating that the results does not depend strongly on the samples. The results for sample II as well as those for sample I are listed in Table 2. However, we would like to note that the results might be changed if we take the other limit of $\beta$; $\beta=1$ corresponding to radial orbits. In this case the best $M/L$ for sample I was found to be $42_{-7}^{+34}$ ($q=0$), and so the $M/L$ would exceed 50 within 68 \% confidence level. Yet the assumption of $\beta=1$ seems too radical, because in this case the pairs suffer from direct encounters that will probably lead them to mergers. In order for bound pairs to survive for many orbital periods, their orbits must be elliptical at least to some degree, and so $\beta$ cannot be too close to unity. On the other hand, perfectly circular orbits ($\beta =-\infty$), are also unlikely because this requires fine tuning of orbital parameters. Therefore, the results with $\beta=0$ presumably represent best the mass-to-light ratio of true pairs. \subsection{$M/L$ for pure spiral pairs} In the $M/L$ determination above, we did not consider the variation of $M/L$ among the sample galaxies. The $M/L$ of galaxies are, however, usually considered to vary depending on the type of galaxies. Indeed, previous binary galaxy studies claimed larger $M/L$ for ellipticals than that of spirals (e.g., Schweizer 1987). Therefore, it is interesting to study the $M/L$ ratio of galaxies for a specific type. The 57 pairs in sample I consist of spirals, S0s, ellipticals and some others such as peculiars. The dominant type among them is spirals, which occupies a fraction of 70\% in sample I, and hence we try to estimate $M/L$ for spiral galaxies. In particular we concentrate on {`}pure{'} spiral pairs which consists of two spiral galaxies later than Sa, because a pair of a spiral and an S0 or elliptical do not necessarily reflect the $M/L$ of spiral galaxies. We selected 30 pure spiral pairs out of 57 pairs in sample I (hereafter sample III), and performed the same analysis described above. The likelihood contours for no clustering case ($q=0$) and for clustering case ($q=1.8$) are plotted in figure 5 and 6, respectively. Most likely value for no clustering case is found to be $15_{-3}^{+5}$ for $\beta = 0$, and $12_{-3}^{+4}$ for $\beta=-\infty$, which are compared to the results for the 57 pairs with mixed types, 35 for $\beta=0$ or 28 for $\beta=-\infty$. The best estimates of $M/L$ for clustering case are similar to those for no clustering case whereas the true pair fraction $f$ is relatively smaller (see table 2). In both cases the difference in $M/L$ between sample I and III is significant, being above the 3$\sigma$ level. Therefore, we conclude the difference is real, and that $M/L$ for spirals are smaller to ellipticals or S0s. This is consistent with previous studies of binary galaxies, although the $M/L$ obtained here are somewhat smaller than those from previous studies. \section{Discussion} \subsection{$M/L$ Dependence on Separation} In the previous sections, pairs are approximated as two point masses. However, real galaxies have finite size, and it could be as much as 100 kpc if dark halos extend well beyond the optical disks. In this case, the approximation of point masses is not valid, and $M/L$ obtained above could be underestimation, particularly for pairs with small separations. Here we investigate the dependency of $M/L$ on the separation of pairs, and test if the assumption of point masses is reasonable for the present samples. We divided 57 pairs in sample I into 3 subgroups depending on the separation. The three subgroups consist of 27 pairs with $0 <R_{\rm p}< 100$ kpc, 12 with $100 <R_{\rm p}< 200$ kpc, and 18 with $200 <R_{\rm p}$ kpc, respectively. The $M/L$ ratios for three subsample were obtained in the same manner described in the previous section. Assuming $\beta = 0$ and $q=0$, we obtained the best estimates of $M/L$ with $1\sigma$ errors to be $36_{-5}^{+10}$, $37_{-17}^{+31}$, and $25_{-13}^{+29}$, respectively. The error bars are increased compared to the results in Section 3 because the number of galaxies in a sample is reduced. figure 7 shows the $M/L$ dependence on the mean separations of pairs. The figure demonstrates that the $M/L$ ratio is almost constant, and that the variations are within the $1\sigma$ error bars. If the density distribution of dark halo is proportional to $r^{-2}$ at a large radius, the $M/L$ increases linearly with radius, and if it is proportional to $r^{-3}$ as suggested by recent simulations (Navarro et al.1996), $M/L$ increases with radius logarithmically. However, figure 7 shows no tendency of increasing $M/L$ at large radii. Therefore, the halos of galaxies as luminous as the Milky Way Galaxies may be truncated within 100 kpc. The indication of constant $M/L$ beyond 100 kpc is consistent with previous studies of binary galaxies. According to Peterson (1979b), the $M/L$ ratio is gradually increasing with radius at $R<100$ kpc, but remains constant beyond it. Schweizer (1987) also showed that $M/L$ does not increase beyond 100 kpc. The $M/L$ estimate for spiral galaxies is also consistent with previous studies. Schweizer (1987) obtained $M/L$ of $21\pm 5$ (V band, absorption corrected) with the sample whose mean separation is about 90 kpc. Peterson (1979b) obtained spiral galaxies{'} $M/L$ of $35\pm 13$ ($H_0 = 50$ km s$^{-1}$ Mpc$^{-1}$) based on 39 pairs with their mean separation of 110 kpc, and Turner (1976b) also obtained $M/L$ for spiral galaxies of $\sim$ 35. Note that in the 70's the correction for the galactic and internal absorption were not usually made, and this partly explains smaller $M/L$ in the present paper. The mean of absorption correction in our sample galaxies is about 0.4 mag, which reduces $M/L$ by $\sim$ 30 \%. Therefore, if similar amount of absorption correction were made, the studies by Peterson (1979b) or Turner (1976b) would give $M/L$ of 23$\sim$27, which are close to our results of $M/L$. Although the $M/L$ obtained in the present paper is not significantly different from previous studies, we would like to emphasize that the mean of absolute separations (not the luminosity corrected separation $R_{\rm p}$) of sample III in this paper is $\sim$ 206 kpc, which is almost twice of those in previous studies. Nevertheless, the $M/L$ ratio does not show any tendency of increase with increasing the separation when compared with previous studies. \subsection{Dark Halo Extent of Spiral Galaxies} Since the mass of spiral galaxies{'} optical disk can be estimated from rotation curves, we can compare the total $M/L$ with optical disk $M/L$, and discuss the extent of dark halos of spiral galaxies. We define $(M/L)_{R25}$ as the ratio of the enclosed mass within $R_{25}$ to the total luminosity, where $R_{25}$ is a radius at which the surface brightness becomes 25 mag per arcsec$^2$. The enclosed mass can be estimated from the HI velocity line width $W_{\rm HI}$. Assuming that $W_{\rm HI}$ corresponds to twice of the rotation velocity, we obtain \begin{equation} (M/L)_{R25} = \frac{R_{25} W_{\rm HI}^2}{4 G L}. \end{equation} The central surface luminosity of spiral galaxies is constant, about 22 mag per arcsec$^2$ (Freeman 1970). If this applies to the spiral galaxies in the binary sample, $R_{25}$ corresponds to about 3 times disk scale length $d$, and $(M/L)_{R25}$ roughly approximates the mass-to-right ratio of the disk. We calculated $(M/L)_{R25}$ of spiral galaxies in sample I for which $R_{25}$ and $W_{\rm HI}$ are available. $R_{25}$ and $W_{\rm HI}$ were taken from de Vaucouleurs et al.(1991), and Huchtmeier and Richter (1989), respectively. The values of $(M/L)_{R25}$ range from 2 to 18 with the average of 7. Note that this $M/L$ is consistent with the previous studies for disk $M/L$; for example, Faber \& Gallagher (1979) obtained the $M/L$ of about 5 ($H_0=50$ km s$^{-1}$ Mpc$^{-1}$). This $M/L$ is compared to the total $M/L$ obtained in the section 3, $M/L$ of 12 $\sim$ 16. The total $M/L$ is somewhat larger than the disk $M/L$, and the difference is almost $3\sigma$ level (see figures 5 and 6). This difference is of course due to the dark halo, and this indicates that under the maximum disk assumption the contributions of dark halo and the optical disk to the total mass of galaxies are comparable. However, the assumption of maximum disk is still controversy. If the disk mass is smaller than that indicated by the maximum rotation velocity within the optical disk, the dark halo could dominantly contribute to the total mass. We can also estimate the extent of dark halos by comparing the total $M/L$ with disk $M/L$. If a flat rotation curve is assumed, the value of $M/L$ increases linearly with radius. In this case, the resultant $M/L$ of 15 implies that the typical halo extends $15/7 \approx 2$ times $R_{25}$. If we adopt disk $M/L$ of 5 according to Faber \& Gallagher (1979), the halo extent is $15/5 \approx 3$ times $R_{25}$. Therefore, the typical dark halo size maybe about 6 to 9 times disk scale length, if $R_{25}$ is $\sim 3$ times disk scale length $d$. This is to be compared with the size of optical disk, which is about $4\sim 5$ times disk scale length (van der Kruit \& Searle 1981). Hence, if the rotation curve is perfectly flat out to the radius at which the halo mass distribution is truncated, the halo size may be about 2 times larger than optical disks. This is, of course, not exactly true when the rotation curve is not completely flat, which is preferred by the recent simulations. In this case, the halo size may be somewhat lager, but it cannot exceed a few hundred kpc as indicated by the $M/L$ constancy. The halo size indicated here is somewhat smaller than those in previous studies, but quite consistent with recent investigations. For instance, a number of declining rotation curves, which may be fitted even with a Keplerian, are found recently (e.g., J\"ors\"ater \& van Moorsel 1995; Olling 1996; Honma \& Sofue 1996). Honma \& Sofue (1997) showed that such declining rotation curves are not uncommon by considering the observational uncertainty. These rotation curves studies are generally based on the observation of HI, which are usually observed out to $5\sim 10$ times disk-scale length. Therefore, the fact that the declining part of rotation curves were found is consistent with the present results. \acknowledgments We are grateful to Y. Sofue for his supervision, and to Y. Tutui and J. Koda for fruitful discussion. This work was financially supported by the Japan Society for the Promotion of Science. \clearpage \section*{figure captions} \def\hangindent=1pc \noindent{\hangindent=1pc \noindent} \hangindent=1pc \noindent Figure 1. Distribution of selected pairs in the $R_{\rm p}$-$V_{\rm p}$ phase space. Figure 1a is for the pairs selected with $a=1.5$, and 1b is for the pairs selected with $a=2.5$. The dotted curve in figure 1b corresponds to the constant $(M/L)_{\rm p}$ of 20. \hangindent=1pc \noindent Figure 2. Likelihood contours in the parameter space of $M/L$ and $f$ for sample I. As for the optical pair distribution no clustering is assumed ($q=0$). Solid lines are for $\beta=0$ (isotropic orbits), and dotted lines are for $\beta=-\infty$ (circular orbits). Three contours for each case correspond to 68, 95 and 99\% level, and the crosses denote the peak of the likelihood. \hangindent=1pc \noindent Figure 3. Likelihood contours same to figure 2, but for $q=1.8$ (clustering case for optical pairs). \hangindent=1pc \noindent Figure 4. Simulated distribution of 57 pairs which are to be compared with figure 1b. The distribution is calculated with the best fit parameters obtained in the Section 3. The dotted curve corresponds to $(M/L)_{\rm p}$ of 20. \hangindent=1pc \noindent Figure 5. Likelihood contours same to figure 2 (no clustering case), but for sample III (30 pure spiral pairs). \hangindent=1pc \noindent Figure 6. Likelihood contours same to figure 3 (clustering case), but for sample III (30 pure spiral pairs). \hangindent=1pc \noindent Figure 7. $M/L$ for three subgroups against the mean separation (see text for the sample). The error bar for $R$ denotes the $1\sigma$ deviation of pairs in the sample. \clearpage
\section{Introduction} \label{sec:intro} Recently, the possible formation of diquark condensates in QCD at finite density has been reinvestigated in a series of papers following Refs. (\cite{arw98}; \cite{r+98}). It has been shown that in chiral quark models with a nonperturbative 4-point interaction motivated from instantons (\cite{cd98}) or nonperturbative gluon propagators (\cite{br98}, \cite{brs99}) the anomalous quark pair amplitudes in the color antitriplet channel can be very large: of the order $\approx 100~ {\rm MeV}$. Therefore, in two-flavor QCD, one expects this diquark condensate to dominate the physics at densities beyond the deconfinement/chiral restoration transition and below the critical temperature ($\approx 50~ {\rm MeV} $) for the occurence of this ``color superconductivity'' (2SC) phase. In a three-flavor theory it has been found (\cite{arw99}, \cite{sw99}) that there can exist a color-flavor locked (CFL) phase for not too large strange quark masses (\cite{abr99}) where color superconductivity is complete in the sense that diquark condensation produces a gap for quarks of all three colors and flavors, which is of the same order of magnitude as that in the two-flavor case. The high-density phases of QCD at low temperatures are most relevant for the explanation of phenomena in rotating compact stars - pulsars. Conversely, the physical properties of these objects (as far as they are measured) could constrain our hypotheses about the state of matter at the extremes of densities. In contrast to the situation for the cooling behaviour of compact stars (\cite{bkv99}) where the CFL phase is dramatically different from the 2SC phase, we don't expect qualitative changes of the magnetic field structure between these two phases. Consequently, we will restrict ourselves here to the discussion of the simpler two-flavor theory first. According to Bailin and Love (1984) the magnetic field of pulsars should be expelled from the superconducting interior of the star due to the Meissner effect and decay subsequently within $\approx 10^4$ years. If their arguments would hold in general, the observation of lifetimes for the magnetic field as large as $10^7$ years (\cite{pines}; \cite{bpp69}) would exclude the occurence of an extended superconducting quark matter phase in pulsars. For their estimate, they used a perturbative gluon propagator which yielded a very small pairing gap and they made the assumption of a homogeneous magnetic field. Since both assumptions seem not to be valid in general, we perform a reinvestigation of the question whether presently available knowledge about the lifetime of magnetic fields of pulsars might contradict the occurence of a color superconducting phase of QCD at high densities. The free energy density in the superconducting quark matter phase with $ud$ diquark pairing ($J^P=0^+$ and color antitriplet index $p$) is given by (\cite{bl84}) \begin{eqnarray} f&=&f_n + \alpha d_{p}^*d_{p} + \frac{1}{2}{\beta}(d_{p}^*d_{p})^2 \nonumber\\ && + \gamma(\nabla d_{p}^* +iq\vec{A}d_{p}^*)(\nabla d_{p} -iq\vec{A}d_{p}) +\frac{B^2}{8 \pi}~, \end{eqnarray} where $\vec{B}={\rm rot}\vec{A}$ is the magnetic induction, $q$ the charge of the $ud$ pair and the coefficients of the free energy are given by the following expressions \begin{eqnarray} \alpha&=&\frac{d n}{d E} t~~,\nonumber\\ {\beta}&=& \frac{d n}{d E} \frac{7 \zeta (3)}{8 (\pi k_{\rm B} T_{\rm c})^2}~~, \nonumber\\ \gamma &=& \frac{d n}{d E} \frac{7 \zeta (3)}{48 (\pi k_{\rm B} T_{\rm c})^2} \frac{p_{\rm F}^2}{\mu^2}=\frac{1}{6}\frac{p_{\rm F}^2}{\mu^2}{\beta}~~, \end{eqnarray} where $t=(T-T_{\rm c})/T_{\rm c}$. Here $T_{\rm c}$ is the critical temperature, $p_{\rm F}$ the quark Fermi momentum, $\mu=\sqrt{p_{\rm F}^2+m^2}$ is the chemical potential (in zeroth order with respect to the coupling constant), $d n/d E=\mu p_{\rm F}/\pi^2$. The Ginzburg-Landau equations for relativistic superconducting quarks are obtained in the usual way \begin{eqnarray} 0&=&\alpha d_{p}+\beta(d_{p}^*d_{p})d_{p}+\gamma(-i\nabla-q\vec{A})^2 d_{p}~, \nonumber\\ \vec{j}&=& iq\gamma(d_{p}\nabla d_{p}^*-d_{p}^*\nabla d_{p})- 2q^2\gamma d_{p}d_{p}^*\vec{A}~. \label{gl} \end{eqnarray} In deriving the expression for the current $\vec{j}$ we have also used the Maxwell equation \begin{equation} {\rm rot} \vec{B}= 4\pi \vec{j}~. \end{equation} The first of the Ginzburg-Landau equations (\ref{gl}) has a solution which corresponds to the Meissner effect ($\nabla d_{p}=0,~\vec{A}=0$ inside of the superconductor): \begin{equation} \Delta^2= |d_{p}|^2=-\frac{\alpha}{\beta} = - \frac{8 t (\pi k_{\rm B} T_{\rm c})^2}{7 \zeta(3)}~. \end{equation} For the case of weak fields ($H<H_{c2}$) one obtains the London equation: \begin{equation} \vec{B} +\lambda_{\rm q}^2 {\rm rot~ rot} \vec{B} = 0~~, \end{equation} where $\lambda_{\rm q}$ is the penetration depth of the magnetic field into the superconducting quark condensate. The region of the change of the order parameter $d_{p}$ can also be determined from (\ref{gl}) via \begin{equation} \xi_{\rm q}^2=-\frac{\gamma}{\alpha} =-\frac{7 \zeta(3)}{48 t (\pi k_{\rm B} T_{\rm c})^2} \left(\frac{p_{\rm F}}{\mu}\right)^2~. \end{equation} \begin{figure}[htb] \psfig{figure=kapnjl.eps,width=8.5cm,height=7.5cm,angle=0} \caption{The dependence of diquark energy gap $\Delta$ and Ginzburg-Landau parameter $\kappa$ on chemical potential $\mu$ and density n for a NJL-type quark interaction (\protect\cite{br99}); n$_0=0.16$ fm$^{-3}$ is the nuclear saturation density. The left dashed line denotes the critical chemical potential of the onset of quark superconductivity (the corresponding baryon number densities are given on the upper scale), the right dashed line - the maximal values of chemical potential and density for stable stellar configurations.\label{kappa}} \end{figure} In the diquark condensate phase with a nonperturbative interaction the energy gap is $\Delta \approx 100$ MeV at $\mu \approx 400$ MeV \footnote{within a dynamical confining quark model the diquark gaps can be even larger than this estimate (\cite{br98}).}, see Fig. \ref{kappa}. We obtain for the coherence length $\xi_{\rm q} = 0.8 \times 10^{-13}$ cm. For the penetration depth of the magnetic field we have \begin{equation} \lambda_{\rm q}=\frac{1}{\sqrt{8\pi\gamma} q d_0} =\sqrt{\frac{-3\pi\mu}{4 q^2 t p_{\rm F}^3}} \approx 3.6\times 10^{-12} {\rm cm}~. \end{equation} The thermodynamical critical field $H_{\rm cm}$ that fully destroys the superconducting state in the case of a superconductor of the first kind is given (\cite{bl84}) \begin{equation} H_{\rm cm}^2=\frac{32 \pi \mu p_{\rm F} (k_{\rm B}T_{\rm c}t)^2}{7 \zeta(3)}~. \end{equation} For the parameter values given above the critical field is $H_{\rm cm}\approx 8.7 \times 10^{17}$ G, i.e. by two orders of magnitude larger than in Ref. (\cite{bl84}). The Ginzburg-Landau parameter $\kappa$ which determines the behaviour of the superconductor in an external magnetic field is given by (\cite{bl84}) \begin{equation} \kappa=\frac{\lambda_{\rm q}}{\xi_{\rm q}} =\sqrt{\frac{{\beta}}{8\pi \gamma^2 q^2}} =132 \frac{\Delta}{\mu}\left(\frac{\mu}{p_{\rm F}}\right)^{5/2}~. \end{equation} For values of $\mu\sim p_{\rm F}\sim 400$ MeV/c and $\Delta=100$ MeV we obtain $\kappa = 34$, see also Fig. 1. Therefore the superconducting quark condensate appears as a superconductor of the second kind into which the external magnetic field can penetrate by forming quantized vortex lines in the interval $H_{c1}<H<H_{c2}$. The upper critical field $H_{c2}$ is determined by \begin{equation} H^{\rm q}_{c2}=-\frac{\alpha}{q \gamma} =\frac{6 \Delta^2}{q}\left(\frac{\mu}{p_{\rm F}}\right)^2 \approx 3 \times 10^{19} {\rm G}~. \end{equation} The magnetic flux of the quark vortex lines $\Phi_{\rm q}$ amounts to \begin{equation} \Phi_{\rm q}=\frac{2 \pi \hbar c}{q}= \frac{2 \pi \hbar c}{e/3} = 6\frac{\pi \hbar c}{e} = 6 \Phi_0~~, \end{equation} where $\Phi_0=2\times 10^{-7}$ G cm$^2$ is the quantum of the proton magnetic flux. The lower critical field for the occurence of quark vortex lines is then \begin{eqnarray} H_{c1}^{\rm q}= \frac{\Phi_{\rm q}}{6\pi\lambda_{\rm q}^2} \ln\frac{\lambda_{\rm q}}{\xi_{\rm q}} = 1.8 \times 10^{16} {\rm G}~. \end{eqnarray} Here we have taken into account the spherical shape of the quark core in a neutron star (\cite{ssm84}). Now we can describe the magnetic structure of a superconducting quark condensate in a pulsar and its time evolution. When during the cooling of the protoneutron star with a quark matter core the critical temperature for the transition to the superconducting state is reached in the presence of a magnetic field, then this field remains in the quark phase in the form of quantized vortex lines. At some point in the further rapid cooling of the star due to neutrino emission the neutrons in the hadronic phase (``npe''-phase) of the star become superfluid. Since the basic interaction between isolated protons resembles that of the neutrons, the protons in the hadronic phase become superfluid too. Since the density of protons in "npe"-phase is only few per cent of the neutron density, the protons will pair in $^1$S$_0$ pairing state (\cite{ccy72}; \cite{ao85}; \cite{bcll92}). The neutrons take part in the rotation, forming a lattice of quantized vortex lines. Because of the strong interaction of the neutrons with the protons a part of the superconducting protons will be entrained by the neutrons (\cite{ss80}; \cite{als84}) and create in the region of the neutron vortex a magnetic field of strength $H(r)$ given by (\cite{ss80}; \cite{ssm83}) \begin{equation} {H}(r)=\hat{\nu}_{\rm n}\frac{k \Phi_0}{2\pi\lambda_{\rm p}^2}\ln\frac{b}{r}~, \end{equation} where $b=\sqrt{\pi \hbar/\sqrt{3}m_{\rm n} \Omega}$ is the lattice spacing of the neutron vortex lattice, $k=(m_{\rm p}^*-m_{\rm p})/m_{\rm p}$ is the entrainment coefficient with the effective mass $m_{\rm p}^*$ and the bare mass $m_{\rm p}$ of the protons; $\hat{\nu}_{\rm n}$ is the unit vector in the direction of the vortex axis, $r$ is the distance from the center of the vortex and $\Omega$ is the angular velocity of the rotation of the star. This field, whose magnitude is determined by the rotation of the star, acts as an external field for the non-entrained protons and creates a cluster of proton vortices with the fluxes $\Phi_0$ in the region around the axis of the neutron vortex where $H(r)>H_{c1}^{\rm p}$. The radius of this region, $\delta_{\rm n}$, equals (\cite{ssm84}) \begin{equation} \delta_{\rm n}=b ({\xi_{\rm p}}/{\lambda_{\rm p}})^{\frac{1}{3|k|}}~. \end{equation} For the pulsar Vela PSR 0833-45 with $\Omega=70$ rad~s$^{-1}$ and $b=10^{-3}$ cm, we have $\delta_{\rm n}=10^{-5}$ cm. While the mean magnetic induction in the star due to proton vortex clusters is of the order $10^{12}$ G, the mean magnetic induction within the cluster reaches values of $4 \times 10^{14}$ G (\cite{ss91}; \cite{ss95}). The magnetic field strength $H(r)$ which occurs in the ``npe''-phase is the strength of the external field relative to the superconducting quark condensate. It reaches the maximum value $H(0)$ close to the center of the neutron vortex, i.e. \begin{equation} H(0)=\frac{k \Phi_0}{2 \pi \lambda_{\rm p}^2} \ln \frac{b}{\xi_{\rm n}} \approx 4.7 \times 10^{16} {\rm G}~, \end{equation} for $\lambda_{\rm p}=30$ fm, $k=0.7$ and a coherence length $\xi_{\rm n}= 30$ fm of the neutron. This external field generates quark vortex lines when the condition $H(r)\ge H_{c1}^{\rm q}$ is fulfilled. The radius of this region is \begin{equation} \delta_{\rm q}=b({\xi_{\rm q}}/{\lambda_{\rm q}})^{\frac{2}{k} (\lambda_{\rm p}/\lambda_{\rm q})^2}=4.3 \times 10^{-7} {~\rm cm}~~. \end{equation} This way, the entrainment current generates a strongly inhomogeneous magnetic structure in the quark condensate: the clusters of quark vortex lines with the fluxes $\Phi_{\rm q}$ and radii $\delta_{\rm q}$, the axes of which are the continuation into the quark phase of the axes of the neutron vortex lines. Since $\delta_{\rm q}$ is by two orders of magnitude smaller than $\delta_{\rm n}$, the mean magnetic induction in the clusters of quark vortex lines increases to a value of the order of $ 10^{18}$ G. When the condition for the applicability of the London approximation $H(0)\ll H_{c2}^{\rm q}$ is fulfilled, then one can apply the modified London equation \begin{equation} \vec{B}+ \lambda_{\rm q}^2 {\rm rot ~rot} \vec{B} = \Phi_{\rm q} \hat{\nu}_{\rm q} \sum \delta(\vec{r}-\vec{r}_{\rm q}) \end{equation} for the description of the magnetic structure of the quark condensate. The density of clusters is equal to the density of neutron vortex lines $n_V=2~\Omega/\kappa_{\rm n}$, where $\kappa_{\rm n}=\pi\hbar/m_{\rm n}$ is the quantum of neutron circulation. We note that between the hadronic phase and quark core there is a mixed phase, in which hadrons coexist with a charged lattice of quark droplets (\cite{ng92}). Since the number densities of neutrons and protons in the mixed state are lower than in the hadronic phase, these particles remain superfluid. So neutron vortices and clusters of proton vortex lines continue through the mixed phase. Therefore the magnetic field will pass through it and enter the quark core, see Fig. \ref{vortex}. \begin{figure} \psfig{figure=vortex1.eps,width=8.5cm,height=8cm,angle=0} \caption{Magnetic field structure in the interior of a hybrid star with $M=1.4 M_{\odot}$; $b$ is the radius of the neutron vortex $\delta_{\rm n}= 10^{-5}$ cm is the radius of the proton vortex cluster, and $\delta_{\rm q} = 4.3 \times 10^{-7}$ cm is that of the quark vortex cluster. For details, see text. \label{vortex}} \end{figure} In the case of small diquark gaps of the order of 1 MeV (\cite{bl84}), when the diquark condensate is a superconductor of the first kind, the magnetic field generated in the ``npe''-phase penetrates into the quark matter core in the form of ordinary cylindrical regions (\cite{ssz97}). The radii of these regions will be of the order of $\delta_{\rm q}$ since the thermodynamical critical field $H_{\rm cm}$ is of the same order as the mean magnetic field in the quark cluster. The clusters of quark vortex lines which appear due to the entrainment effect in the ``npe''-phase will interact with those which are formed by the initial magnetic field (fossil field). This interaction obviously implies that quark vortex lines will not be expelled from the quark core of the star within a time scale of $\tau=10^4$ years as suggested in (\cite{bl84}). We note that the evolution of the magnetic field is intimately related to the rotational history of the star. In particular, the magnetic field of the quark core will decay because of the outward motion of neutron vortices when the star spins down. This behavior results from the fact that the magnetic clusters inside the quark core are the continuation of neutron vortices. Therefore the characteristic decay time of the magnetic field for the whole star (and also for quark core) is comparable to the pulsar's slowing down time, which corresponds to the life time of the pulsar. In conclusion, we find that the occurence of a superconducting quark matter core in pulsars does not contradict the observational data which indicate that magnetic fields of pulsars have life times larger than $10^7$ years (\cite{pines}). This holds true for small diquark gaps of the order of $1$ MeV (\cite{bl84}) as well as for larger ones as obtained recently (\cite{arw98}; \cite{r+98}; \cite{cd98}; \cite{br98}) using effective models for the nonperturbative quark-quark interaction. \begin{acknowledgements} K.M.S. and D.M.S. acknowledge the hospitality of the Department of Physics at the University of Rostock where this research has been started. We thank H. Grigorian, K. Rajagopal, G. R\"opke, A.D. Sedrakian and D.N. Voskresensky for their discussions and comments. \end{acknowledgements}
\section{Introduction} The evolution of a star is made of a succession of ``controlled" thermonuclear burning stages interspersed with phases of gravitational contraction. The latter stages are responsible for a temperature increase, the former ones producing nuclear energy and composition changes. As is well known, hydrogen and helium burning in the central regions or in peripheral layers of a star are key nuclear episodes, and leave clear observables, especially in the Hertzsprung-Russell diagram, or in the stellar surface composition. These photospheric abundance signatures may result from so-called ``dredge-up'' phases, which are expected to transport the H- or He-burning ashes from the deep production zones to the more external layers. This type of surface contamination is encountered especially in low- and intermediate-mass stars on their first or asymptotic branches, where two to three dredge-up episodes have been identified by stellar evolution calculations. Nuclear burning ashes may also find their way to the surface of non-exploding stars by rotationally-induced mixing, which has been started to be investigated in some detail (Heger 1998), or by steady stellar winds, which have their most spectacular effects in massive stars of the Wolf-Rayet type. The confrontation between the wealth of observed elemental or isotopic compositions and calculated abundances can provide essential clues on the stellar structure from the main sequence to the red giant phase, and much has indeed been written on this subject. Of course, the information one can extract from such a confrontation is most astrophysically useful if the discussion is freed from nuclear physics uncertainties to the largest possible extent. Thanks to the impressive skill and dedication of some nuclear physicists, remarkable progress has been made over the years in our knowledge of reaction rates at energies which are as close as possible to those of astrophysical relevance (e.g. Rolfs \& Rodney 1988). Despite these efforts, important uncertainties remain. This relates directly to the enormous problems the experiments have to face in this field, especially because the energies of astrophysical interest for charged-particle-induced reactions are much lower than the Coulomb barrier energies. As a consequence, the corresponding cross sections can dive into the nanobarn to picobarn abyss. In general, it has not been possible yet to measure directly such small cross sections. Theoreticians are thus requested to supply reliable extrapolations from the lowest energies attained experimentally to those of most direct astrophysical relevance. Recently, a new major challenge has been taken up by a consortium of European laboratories with the build-up of well documented and evaluated sets of experimental data or theoretical predictions for a large number of astrophysically interesting nuclear reactions (Angulo et al. 1999). This compilation of reaction rates, referred to as NACRE (Nuclear Astrophysics Compilation of REaction rates; see Sect.~2 for some details), comprises in particular the rates for all the charged-particle-induced nuclear reactions involved in the ``cold'' pp-, CNO, NeNa and MgAl chains, the first two burning modes being essential energy producers, all four being important nucleosynthesis agents. It also includes the main reactions involved in non-explosive helium burning. The aim of this paper is to calculate with the help of the NACRE data the abundances of the different isotopes of the elements from C to Al involved in the non-explosive H (Sects.~3 - 5) and He (Sect.~6) burnings, special emphasis being put on the impact of the reported remaining rate uncertainties on the derived abundances. The yields from the considered burning modes are calculated by combining in all possible ways the lower and upper limits of all the relevant reaction rates. One ``reference'' abundance calculation is also performed with all the recommended NACRE rates. Note that the pp-chains are not considered here. A solar neutrino analysis based on preliminary NACRE data for the pp reactions can be found in Castellani et al. (1997). Our extensive abundance uncertainty analysis is performed in the framework of a parametric model assuming that H burning takes place at a constant density $\rho = 100$~g~cm$^{-3}$ and at constant temperatures between $T_6=10$ and 80 ($T_n$ is the temperature in units of $10^n~{\rm K}$). The corresponding typical values adopted for He burning are $\rho = 10^4$~g~cm$^{-3}$ and $T_8=1.5$ and 3.5. These ranges encompass typical burning conditions in a large variety of realistic stellar models. For the study of H-burning, initial abundances are assumed to be solar (Anders \& Grevesse 1989). For He-burning, we adopt the abundances resulting from H burning at $T_6=60$ and $\rho = 100$~g~cm$^{-3}$ calculated with the use of the NACRE recommended rates. The H- and He-burning nucleosynthesis is followed until the H and He mass fractions drop to $10^{-5}$. In spite of its highly simplistic aspect, this analysis provides results that are of reasonable qualitative value, as testified by their confrontation with detailed stellar model predictions. Most significant, these parametric calculations have the virtue of identifying the rate uncertainties whose impact may be of significance on abundance predictions at temperatures of stellar relevance. They thus serve as a guide in the selection of the nuclear uncertainties that have to be duly analyzed in detailed model stars, particularly in order to perform meaningful confrontations between abundance observations and predictions. They are also hoped to help nuclear astrophysicists pinpointing the rate uncertainties that have to be reduced most urgently. \section{The NACRE compilation in a nutshell} A detailed information about the procedure adopted to evaluate each of the NACRE reaction rates and about the derived values can be found in Angulo et al. (1999), or in electronic form at {\it http://astro.ulb.ac.be}, which also offers the possibility of generating interactively tables of reaction rates for networks and temperature grids selected by the user\footnote{This electronic address also provides many other nuclear data of nuclear astrophysics interest}. It is clearly impossible to go here into the details of the NACRE procedure. Let us just emphasize some of its specificities: \noindent (1) For each reaction, the non-resonant and broad-reso\-nan\-ce contributions to its rate are evaluated numerically in order to avoid the approximations which are classically made (see Fowler et al. 1975 for details) in order to allow analytical rate evaluations; \noindent (2) Narrow or subthreshold resonances are in general approximated by Breit-Wigner shapes, and their contributions to the reaction rates are approximated in the usual analytical way (e.g. Fowler et al. 1975). However, in some cases, the resonance data are abundant enough to allow a numerical calculation avoiding these approximations; \noindent (3) For each reaction, NACRE provides a recommended ``adopted'' rate, along with realistic lower and upper limits. The adopted values of, and the limits on the resonance contributions are derived from weighted averages duly taking into account the uncertainties on individual measurements, as well as the different measurements that are sometimes available for a given resonance [see Eq.~(15) of Angulo et al. 1999]. For non-resonant contributions, $\chi^2$-fits to available data provide the recommended values along with the lower and upper limits, as the experimental uncertainties on one set of data and the differences between various sets, if available, are taken into account in the $\chi^2$-procedure. It is worth stressing at this point that enough information is provided by NACRE for helping the user to tailor his own preferred rates if he wants. The procedure just sketched in (1) - (3) is the selected standard methodology, and has the advantage of being easily reproducible and of avoiding any subjective renormalization of different experimental data sets.Quite clearly, however, the large variety of different situations makes unavoidable some slight modifications of the standard procedure in some cases. These specific adjustments are clearly identified and discussed in Angulo et al. (1999); \begin{figure} \resizebox{\hsize}{!}{\includegraphics[scale=0.2]{H1352.f1}} \caption{Reactions of the CNO cycles. The dashed line represents the possible leakage out of the cycles } \label{fig01} \end{figure} \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[angle=-90]{H1352.f2}} \hfill \vspace{-5cm} \caption{ {\it Left and right panels:} Time variations of the mass fractions of the stable C and N isotopes versus the amount of hydrogen burned at constant density $\rho=100\;{\rm g/cm^3}$ and constant temperatures $T_6=25$ and 55. The H mass fraction is noted $X$(H), the subscript 0 corresponding to its initial value; {\it Middle panel:} Mass fractions of the same nuclides at H exhaustion [X(H)=$10^{-5}$] as a function of $T_6$. The shaded areas delineate the uncertainties resulting from the reaction rates } \label{Fig:CNOyields} \end{figure*} \noindent (4) A theoretical (Hauser-Feshbach) evaluation of the contribution to each rate of the thermally populated excited states of the target is also provided. It has to be noted that the widely used compilation of Caughlan \& Fowler (1988, hereafter referred to as CF88) provides uncertainties for some rates only, while the contribution of excited target states is derived in most cases from a rough (referred to as ``equal strength'') approximation; \noindent (5) It has to be emphasized that the major goal of the NACRE compilation is to provide numerical reaction rates in tabular form (see {\it http://astro.ulb.ac.be}). This philosophy differs markedly from the one promoted by the previous widely used compilations (CF88, and references there\-in), and is expected to lead to more accurate rate data. However, for completeness, NACRE also provides analytical approximations (Angulo et al. 1999) that differ in several respects from the classically used expressions (CF88, and references therein). \section{The CNO Cycles} \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[angle=-90]{H1352.f3}} \hfill \caption{Same as Fig.~\ref{Fig:CNOyields}, but for the O and F nuclides} \label{Fig:OFyields} \end{figure*} The reactions of the CNO cycles are displayed in Fig.~1. As is well known, their net result is the production of \chem{He}{4} from H, and the transformation of the C, N and O isotopes mostly into \chem{N}{14} as a result of the relative slowness of \reac{N}{14}{p}{\gamma}{O}{15} with respect to the other involved reactions. This \chem{N}{14} build-up is clearly seen in Fig.~2. As shown in Fig.~1, three nuclides are important branching points for the CNO cycles. The first one is \chem{N}{15}. At $T_6=25$, \reac{N}{15}{p}{\alpha}{C}{12} is 1000 times faster than \reac{N}{15}{p}{\gamma}{O}{16}, and the CN cycle reaches equilibrium already before $10^{-3}$ of the initial protons have been burned. The second branching is at \chem{O}{17}. The competing reactions \reac{O}{17}{p}{\alpha}{N}{14} and \reac{O}{17}{p}{\gamma}{F}{18} determine the relative importance of cycle II over cycle III (Fig.~\ref{fig01}). The uncertainties on these rates have been strongly reduced in the last years. The rate of \reac{O}{17}{p}{\alpha}{N}{14} recommended by NACRE is larger than the CF88 one by factors of 13 and 90 at $T_6=20$ and 80, respectively. Smaller deviations, though reaching a factor of 9 at $T_6=50$, are found for the \reac{O}{17}{p}{\gamma}{F}{18} rate. \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[scale=0.30,angle=-90]{H1352.f4}} \caption{ Bottom panels: Temperature dependence of Maxwellian-averaged reaction rates (expressed in cm$^3$ mol$^{-1}$ s$^{-1}$) from NACRE for proton capture by \chem{O}{17} (left panel) and by \chem{O}{18} and \chem{F}{19} (right panel). The rate uncertainties given by NACRE are represented by the shaded area. Top panels: Ratio between the NACRE and CF88 Maxwellian-averaged reaction rates} \label{Fig:O17pag} \end{figure*} The oxygen isotopic composition is shown in Fig.~3. As it is well known, it depends drastically on the burning temperature. In particular, \chem{O}{17} is produced at $T_6 \lsimeq 25$, but is destroyed at higher temperatures. This has the important consequence that the amount of \chem{O}{17} emerging from the CNO cycles and eventually transported to the stellar surface is a steep function of the stellar mass. This conclusion could get some support from the observation of a large spread in the oxygen isotopic ratios at the surface of red giant stars of somewhat different masses (Dearborn 1992, and references therein). Fig.~3 also demonstrates that the oxygen isotopic composition cannot be fully reliably predicted yet at a given temperature as a result of the cumulative uncertainties associated with the different production and destruction rates. Finally, the leakage from cycle III is determined by the ratio of the \reac{O}{18}{p}{\gamma}{F}{19} and \reac{O}{18}{p}{\alpha}{N}{15} rates (Fig.~\ref{fig01}). At the temperatures of relevance, \reac{O}{18}{p}{\gamma}{F}{19} is roughly 1000 times slower than \reac{O}{18}{p}{\alpha}{N}{15}, in relatively good agreement with CF88 (Fig.~\ref{Fig:O17pag}), undermining the path leading to the production of \chem{F}{19}. However, at low temperatures, large uncertainties still affect the \reac{O}{18}{p}{\gamma}{F}{19} rate. In fact, its upper bound could be comparable to the \reac{O}{18}{p}{\alpha}{N}{15} rate, and at the same time larger than the \reac{F}{19}{p}{\alpha}{O}{16} rate at $T_6 \lsimeq 20$. As a result, some \chem{F}{19} might be produced, in contradiction with the conclusion drawn from the adoption of the CF88 rates. Fig.~3 indeed confirms that fluorine could be overproduced (with respect to solar) by up to a factor of 100 at H exhaustion when $T_6 \approx 15$. However, Fig. 3 also reveals that the maximum \chem{F}{19} yields that can be attained remain very poorly predictable as a result of the rate uncertainties. In fact, some hint of a non-negligible production of fluorine by the CNO cycles might come from the observation of fluorine abundances slightly larger than solar at the surface of red giant stars considered to be in their post-first dredge-up phase (Jorissen et al. 1992; Mowlavi et al. 1996). As far as \reac{O}{18}{p}{\alpha}{N}{15} is concerned, let us also mention that Huss et al. (1997) have speculated that its rate could be about 1000 times larger than the one adopted by CF88 and NACRE at temperatures of about $15\times 10^6$ K. This proposal has been made in order to explain the N isotopic composition measured in some presolar grains. It is clearly fully incompatible with the NACRE analysis. Finally, let us note that \reac{F}{19}{p}{\alpha}{O}{16} is always much faster than \reac{F}{19}{p}{\gamma}{Ne}{20}. Any important leakage out of the CNO cycles to \chem{Ne}{20} is thus prevented, this conclusion being independent of the remaining rate uncertainties. \section{The NeNa Chain} \begin{figure}[h] \resizebox{\hsize}{!}{\includegraphics[scale=0.34]{H1352.f5}} \caption[]{ \label{Fig:NeNa} Same as Fig.~1, but for the NeNa and MgAl chains} \end{figure} \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[scale=0.30,angle=-90]{H1352.f6}} \caption{\label{Fig:Ne21Ne22pg} Same as Fig.~\ref{Fig:O17pag}, but for \reac{Ne}{20}{p}{\gamma}{Na}{21}, \reac{Ne}{21}{p}{\gamma}{Na}{22}, \reac{Ne}{22}{p}{\gamma}{Na}{23} (left panel) and \reac{Na}{23}{p}{\gamma}{Mg}{24}, \reac{Na}{23}{p}{\alpha}{Ne}{20} (right panel) } \end{figure*} The NeNa chain is illustrated in Fig.~\ref{Fig:NeNa}, while Fig.~\ref{Fig:Ne21Ne22pg} displays some relevant NACRE reaction rates, and their, sometimes quite large, uncertainties. These affect in particular the proton captures by \chem{Ne}{21}, \chem{Ne}{22} and \chem{Na}{23}. In contrast, the \reac{Ne}{20}{p}{\gamma}{Na}{21} rate may be considered as relatively well determined. Some of these rates may also deviate strongly from the CF88 proposed values. \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[angle=-90]{H1352.f7}} \hfill \caption[~NeNa yields] {Same as Fig.~\ref{Fig:CNOyields}, but for the nuclides involved in the NeNa chain} \label{Fig:NeNayields} \end{figure*} The NACRE rates are used to compute the abundances shown in Fig.~\ref{Fig:NeNayields}. A slight alteration of the initial \chem{Ne}{20} abundance is visible only for $T_6 \gsimeq$ 50. However, an unnoticeable \chem{Ne}{20} destruction is sufficient to lead to a significant increase of the abundance of the rare \chem{Ne}{21} isotope through \reacbp{Ne}{20}{p}{\gamma}{Na}{21}{Ne}{21} at $T_6 \lsimeq 40$. At higher temperatures, \reacbp{Ne}{21}{p}{\gamma}{Na}{22}{Ne}{22} destroys \chem{Ne}{21}. As a result, the \chem{Ne}{21} abundance at H exhaustion is maximum when H burns in the approximate $30 \lsimeq T_6 \lsimeq 35$ range. This conclusion may, however, be altered if the upper limit of the \reac{Ne}{21}{p}{\gamma}{Na}{22} rate is adopted instead. The \chem{Na}{23} yield has raised much interest recently, following the discovery at the surface of globular cluster red giant stars of moderate sodium overabundances which correlate or anti-correlate with the amount of other elements (like C, N, O, Mg and Al) also involved in cold H burning (Denissenkov et al. 1998; Kraft et al. 1998, and references therein). This situation may be the signature of the dredge-up to the stellar surface of the ashes of the NeNa chain. The \chem{Na}{23} production results from \reac{Ne}{22}{p}{\gamma}{Na}{23}, while \reac{Na}{23}{p}{\gamma}{Mg}{24} and \reac{Na}{23}{p}{\alpha}{Ne}{20} are responsible for its destruction, which can be substantial at $T_6 \gsimeq 60$. Unfortunately, our knowledge of these three reaction rates remains very poor, with uncertainties that can amount to factors of about 100 to $10^4$ in certain temperature ranges (see Fig.~\ref{Fig:Ne21Ne22pg}). As indicated in Fig.~\ref{Fig:NeNayields}, this situation prevents an accurate prediction of the \chem{Na}{23} yields when $T_6 \gsimeq 50$. More precisely, the spread in the \chem{Na}{23} abundance at H exhaustion reaches a factor of 100 at these temperatures. The possible cycling character of the NeNa chain is determined by the ratio of the rates of \reac{Na}{23}{p}{\alpha}{Ne}{20} and of \reac{Na}{23}{p}{\gamma}{Mg}{24}. Fig.~\ref{Fig:Ne21Ne22pg} indicates that the former reaction is predicted to be faster than the latter one at $T_6 \lsimeq 50$ only. In this case, the NeNa chain is indeed a cycle. However, at higher temperatures, an important leakage to the MgAl chain can be expected, unless future experiments confirm the lower bound of the uncertain \reac{Na}{23}{p}{\gamma}{Mg}{24} rate. \section{The MgAl Chain} The MgAl chain is depicted in Fig.~5. It involves in particular \chem{Al}{26}. Its long-lived ($t_{1/2} =$ \ten{7.05}{5} y) \chem{Al^g}{26} ground state and its short-lived ($t_{1/2} = 6.35$ s) \chem{Al^m}{26} isomeric state are out of thermal equilibrium at the temperatures of relevance for the non-explosive burning of hydrogen (Coc \& Porquet 1998). They have thus to be considered as separate species in abundance calculations. The status of our present knowledge of some important reactions of the MgAl chain is depicted in Fig.~\ref{Fig:Alg26pg}, while the yield predictions for the species involved in this chain are presented in Fig.~\ref{Fig:MgAlyields}. Let us first discuss the situation resulting from the use of the NACRE adopted rates. The most abundant nuclide is \chem{Mg}{24}, the concentration of which remains unaffected, at least for $T_6 \lsimeq 60$. In contrast, \chem{Mg}{25} is significantly transformed by proton captures into \chem{Al^g}{26} at $T_6 \gsimeq 30$. At $T_6 \gsimeq 50$, the leakage from the NeNa cycle starts affecting the MgAl nucleosynthesis through a slight increase of the \chem{Mg}{24} abundance, followed by a modest enhancement of the \chem{Mg}{25}, \chem{Al^g}{26} and \chem{Al}{27} concentrations (Fig.~\ref{Fig:MgAlyields}). At temperatures $T_6 \gsimeq 70$, the \chem{Mg}{24} accumulation starts turning into a depletion by proton captures, which contributes to a further increase in the \chem{Mg}{25}, \chem{Al^g}{26} and \chem{Al}{27} abundances. This build-up cannot be significantly hampered by the destruction of these species by proton captures, as a result of their relative slowness. Among these reactions, \reac{Al}{27}{p}{\alpha}{Mg}{24} and \reac{Al}{27}{p}{\gamma}{Si}{28} are of special interest, as the ratio of their rates determines in particular the leakage out of the MgAl chain. The adopted NACRE rate of the former reaction is 20 to 100 times slower than the CF88 one in the considered temperature range, and turns out to be slower than the (p,$\gamma$) channel for $T_6 \gsimeq 60$ (Fig.~\ref{Fig:Alg26pg}), so that no cycling back is possible in these conditions. \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[scale=0.30,angle=-90]{H1352.f8}} \caption{ \label{Fig:Alg26pg} Same as Fig.~\ref{Fig:O17pag}, but for \reac{Mg}{26}{p}{\gamma}{Al}{27}, \reac{Al^g}{26}{p}{\gamma}{Al}{27} (left panel) and \reac{Al}{27}{p}{\gamma}{Si}{28}, \reac{Al}{27}{p}{\alpha}{Mg}{24} (right panel) } \end{figure*} \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[angle=-90]{H1352.f9}} \hfill \caption {Same as Fig. \ref{Fig:CNOyields}, but for the nuclides involved in the MgAl chain} \label{Fig:MgAlyields} \end{figure*} It is noticeable that the \chem{Mg}{26} abundance at H exhaustion is almost temperature independent. This trend differs from the behaviour of the concentrations of the other Mg and Al isotopes, and results from two factors. First, the adopted \chem{Mg}{26} proton capture is slow enough (about ten times slower than prescribed by CF88) for preventing \chem{Mg}{26} to be destroyed at the considered temperatures. Second, \chem{Mg}{26} is bypassed by the nuclear flow associated with the leakage from the NeNa chain at $T_6 \gsimeq 50$. The reaction \reac{Al^g}{26}{p}{\gamma}{Al}{27} is indeed predicted to be faster than the \chem{Al^g}{26} $\beta$-decay in this temperature domain. Various aspects of the above analysis may be affected by remaining rate uncertainties. In fact, the only proton captures whose rates are now put on safe grounds are \reac{Mg}{24}{p}{\gamma}{Al}{25} (for which NACRE and CF88 are in good agreement) and \reac{Mg}{25}{p}{\gamma}{Al}{26} (for which the NACRE adopted rate is about 5 times slower than the CF88 one at $T_6 < 80$). In spite of much recent effort, the other proton capture rates of the MgAl chain still show more or less large uncertainties in the considered temperature range, as illustrated in Fig.~\ref{Fig:Alg26pg}. Due consideration of these uncertainties indicates in particular (see Fig.~\ref{Fig:MgAlyields}) that, for $T_6 \gsimeq 50$, \chem{Mg}{24} could be more strongly destroyed than stated above, while \chem{Mg}{26} could be substantially transformed into \chem{Al}{27} if the NACRE upper limits on the \reac{Mg}{24}{p}{\gamma}{Al}{25} and \reac{Mg}{26}{p}{\gamma}{Al}{27} rates were selected. It is also important to note that the abundances at H exhaustion of \chem{Al^g}{26} and \chem{Al}{27} are not drastically affected by the uncertainties left in their proton capture rates, even if these uncertainties can be quite large (for example, the \reac{Al^g}{26}{p}{\gamma}{Si}{27} rate is uncertain by more than a factor of $10^3$ at $T_6 \gsimeq 50$). This situation results from the fact that even the highest NACRE proton capture rates are not fast enough for leading to a substantial destruction of the two Al isotopes by the time H is consumed\footnote{Arnould et al. (1995) have reached a different conclusion due to a trivial mistake in the \reac{Al^g}{26}{p}{\gamma}{Si}{27} rate used in their calculations}. In contrast, the exact conditions under which the MgAl chain is cycling cannot be reliably specified yet in view of the large uncertainties still affecting the \reac{Al}{27}{p}{\alpha}{Mg}{24} and \reac{Al}{27}{p}{\gamma}{Si}{28} rates. The possibility for the MgAl chain to produce substantial amounts of \chem{Al^g}{26} is of high interest in view of the prime importance of this radionuclide in cosmochemistry and $\gamma$-ray line astronomy. There is now ample observational evidence that \chem{Al}{26} has been injected live in the forming solar system before its in situ decay in various meteoritic inclusions (MacPherson et al. 1995). Its presence in extinct form is also demonstrated in various types of presolar grains of supposedly circumstellar origin identified in primitive meteorites (e.g. Zinner 1995). The present-day galactic plane also contains \chem{Al^g}{26}, as shown by the observation of a 1.8 MeV $\gamma$-ray line associated with its $\beta$-decay (e.g. Prantzos \& Diehl 1996). The MgAl chain has also a direct bearing on the puzzling Mg-Al anticorrelation observed in globular cluster red giants. Denissenkov et al. (1998) have speculated that a strong low-energy resonance could dominate the rate of \reac{Mg}{24}{p}{\gamma}{Al}{25} at typical cold H-burning temperatures, and could help explaining these observations. There is at present no support of any sort to such a resonant enhancement of this rate. \section{Helium burning} The NACRE compilation also provides recommended rates and their lower and upper limits for most of the $\alpha$-captures involved in the non-explosive burning of helium. The impact of the remaining rate uncertainties on the abundances of the elements up to Al affected by He burning is evaluated in our parametric model for two sets of conditions: {\it (i)} $\rho = 10^4$~g~cm$^{-3}$ and $T_8=1.5$, adopted to characterize the central or shell He-burning phases of intermediate-mass stars ($M\simeq 6$~M$_{\odot}$), and {\it (ii)} $\rho = 10^4$~g~cm$^{-3}$ and $T_8=3.5$, which can be encountered at the end of the He burning phase in the core of massive stars or in AGB thermal pulses. The initial abundances used in these calculations are adopted as described in Sect.~1. In contrast to the H-burning case, the abundances during He burning exhibit some sensitivity to density, as it enters differently the $3\alpha$ reaction rate and the other $\alpha$-capture rates. Consequently, the results presented here should not be used to infer abundances resulting from He burning in specific stellar models, where the time evolution of the temperature and the density may play an important role on the final He-burning composition. It has also to be noted that the neutrons produced by \reac{C}{13}{\alpha}{n}{O}{16} or \reac{Ne}{22}{\alpha}{n}{Mg}{25} during He burning lead us to extend the nuclear network to all (about 500) the s-process nuclides up to Bi. Figs.~\ref{fig_he1} and \ref{fig_he2} illustrate the evolution during He burning in the two situations mentioned above of the abundances of all the stable nuclides between $^{12}{\rm C}$ and $^{27}{\rm Al}$ (plus $^{26}{\rm Al}$). At low temperature ($T_8 \approx 1.5$; Figs.~\ref{fig_he1}a and \ref{fig_he2}a), the main reaction flows are \noindent a) $2\alpha(\alpha,\gamma)^{12}{\rm C}$, followed by \reac{C}{12}{\alpha}{\gamma}{O}{16} at the very end of He burning. The factor of 2 uncertainty in the rate of \reac{C}{12}{\alpha}{\gamma}{O}{16} (Fig.~\ref{Fig:rateHe}) is responsible for the error bars on the $^{16}{\rm O}$ abundance; \noindent b) $^{14}{\rm N}(\alpha,\gamma)^{18}{\rm F}(\beta^+)^{18}{\rm O}$, followed by $^{18}{\rm O}(\alpha,\gamma)^{22}{\rm Ne}$ at the end of He burning. The resulting $^{22}{\rm Ne}$ does not burn at the considered low temperature\footnote{In detailed stellar models, the temperature increases to values in excess of $T_8=3$ towards the end of core (or shell) He-burning. This may lead to the destruction of $^{22}{\rm Ne}$ by $(\alpha,n)$ (with a concomitant production of neutrons) or $(\alpha,\gamma)$ reactions, as illustrated on Fig.~\ref{fig_he2}b}. The uncertainties of a factor of 1.5 and 5 at $T_8=1.5$ in the NACRE rates of $^{14}{\rm N}(\alpha,\gamma)^{18}{\rm F}$ and $^{18}{\rm O}(\alpha,\gamma)^{22}{\rm Ne}$, respectively (Fig.~\ref{Fig:rateHe}), are responsible for the wide range of predicted $^{18}{\rm O}$ and $^{22}{\rm Ne}$ abundances. A much larger $^{18}{\rm O}$ abundance at the end of He burning would result if use were made of the CF88 rate, which is about 220 times smaller than the NACRE one (Fig.~\ref{Fig:rateHe}). \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[angle=-90]{H1352.f10}} \hfill \caption{ Mass fractions of the stable C to F isotopes versus the amount of $^4$He burned at constant density $\rho=10^4\;{\rm g/cm^3}$ and constant temperature $T_8=1.5$ (a: left panel) or $T_8=3.5$ (b: right panel). The $^4$He mass fraction is denoted $X$(He), the subscript 0 corresponding to its initial value} \label{fig_he1} \end{figure*} \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[angle=-90]{H1352.f11}} \hfill \caption{Same as Fig.~\ref{fig_he1} for the nuclides from Ne to Al} \label{fig_he2} \end{figure*} \begin{figure*} \hspace*{3cm} \resizebox{13cm}{!}{\includegraphics[angle=-90]{H1352.f12}} \hfill \caption{Same as Fig.~\ref{Fig:O17pag} for some $\alpha$-capture reactions } \label{Fig:rateHe} \end{figure*} \begin{figure}[h] \resizebox{\hsize}{!}{\includegraphics[scale=0.30]{H1352.f13}} \caption{ Neutron density versus the amount of helium burned at $\rho=10^4\;{\rm g\ cm}^{-3}$ and $T_8=1.5$ (solid line) or $T_8=3.5$ (dashed line). The initial \chem{C}{13} mass fraction is adopted equal to $10^{-4}$, which is obtained at the end of the CNO cycle operating at $T_6 = 60$ and $\rho = 100$ g cm$^{-3}$ (Sect.~3)} \label{fig_he3} \end{figure} The neutron density resulting from \reac{C}{13}{\alpha}{n}{O}{16} is shown in Fig.~\ref{fig_he3}, along with its associated uncertainty. Albeit small, this neutron irradiation is responsible for the $^{15}{\rm N}$ and $^{19}{\rm F}$ abundance peaks seen in Fig.~\ref{fig_he1}a. They result from $^{14}{\rm N}(\alpha,\gamma)^{18}{\rm F}(\beta^+)^{18}{\rm O}({\rm p},\alpha)^{15}{\rm N}(\alpha,\gamma)^{19}{\rm F}$, the protons originating from $^{14}{\rm N}({\rm n,p})^{14}{\rm C}$. Towards the end of He burning, $^{19}{\rm F}$ is destroyed by $^{19}{\rm F}(\alpha,{\rm p})^{22}{\rm Ne}$. Shell He burning in AGB stars or central He burning in Wolf-Rayet stars have been proposed as a major site for the galactic production of $^{19}{\rm F}$ (Goriely et al. 1989; Meynet \& Arnould 1996, 1999; Mowlavi et al. 1998). For AGB stars, these predictions have been confirmed by the observation of $^{19}{\rm F}$ overabundances in some of these objects (Jorissen et al. 1992). Incomplete He-burning (e.g. in Wolf-Rayet stars) may also contribute to the galactic enrichment in primary $^{15}{\rm N}$, as required by the observations of this nuclide in the interstellar medium (G\"usten \& Ungerechts 1985). The large \chem{Al}{26} abundance seen on Fig.~\ref{fig_he2}a results from the particular choice of initial conditions (see Sect.~1), since \chem{Al}{26} is not produced in the conditions prevailing during He-burning. Its rapid drop close to the end of He burning results from the combined effect of $\beta$-decay and $^{26}{\rm Al}({\rm n,p})^{26}{\rm Mg}$ making use of the few neutrons liberated by $^{22}{\rm Ne}(\alpha, {\rm n})^{25}{\rm Mg}$. At higher temperatures (Figs.~\ref{fig_he1}b and \ref{fig_he2}b), the He-burning nucleosynthesis of the elements up to about Al is essentially the same as in the low temperature case. The major differences are observed for \chem{O}{18}, \chem{F}{19}, \chem{Ne}{21}, \chem{Ne}{22}, \chem{Mg}{25}, \chem{Mg}{26} and \chem{Al}{26}, and are mainly due to a larger neutron production by \reac{C}{13}{\alpha}{n}{O}{16}, \reac{O}{18}{\alpha}{n}{Ne}{21} and \reac{Ne}{22}{\alpha}{n}{Mg}{25}. Note that \reac{O}{18}{\alpha}{n}{Ne}{21} is about 150 times slower than $^{18}{\rm O}(\alpha,\gamma) ^{22}{\rm Ne}$ in these conditions, but is fast enough to keep the neutron density above $N_{\rm n}=10^9\; {\rm cm}^{-3}$ (Fig.~\ref{fig_he3}). These neutrons allow protons to be produced by the reactions $^{14}{\rm N(n,p)}^{14}{\rm C}$ and $^{18}{\rm F}({\rm n,p})^{18}{\rm O}$. Additional protons come from $^{18}{\rm F}(\alpha,{\rm p})^{21}{\rm Ne}$. As a result, $^{15}{\rm N}$ is produced via $^{18}{\rm F}({\rm n},\alpha)^{15}{\rm N}$, $^{18}{\rm O}({\rm p},\alpha)^{15}{\rm N}$,\\ $^{14}{\rm N}({\rm p},\gamma)^{15}{\rm O}(\beta^+)^{15}{\rm N}$ and $^{18}{\rm F}({\rm p},\alpha)^{15}{\rm O}(\beta^+)^{15}{\rm N}$. The production of $^{19}{\rm F}$ follows from $^{15}{\rm N}(\alpha,\gamma)^{19}{\rm F}$. Since most of the involved reactions have better known rates at $T_8=3.5$ than at $T_8=1.5$, the corresponding error bars on the abundances are smaller at higher temperature. Neutrons are also responsible for the destruction of any \chem{Al}{26} that may survive the former H-burning episode. The operation of $^{22}{\rm Ne}(\alpha,{\rm n})^{25}$Mg at the end of He burning leads to a non-negligible neutron irradiation which triggers a weak s-process leading to the overproduction of the $70 \lsimeq A \lsimeq 90$ s-nuclei. Unfortunately, the rate of $^{22}{\rm Ne}(\alpha,{\rm n})^{25}$Mg remains quite uncertain (Fig.~\ref{Fig:rateHe}), even at temperatures as high as $T_8=3.5$ (in this case by a factor of 25). The resulting uncertainty on the neutron density amounts to a factor of 10 (Fig.~\ref{fig_he3}), while the total neutron exposure spans the range 0.1 -- 0.3 mbarn$^{-1}$. Finally, the $\alpha$-captures by the Ne isotopes are fast enough at temperatures $T_8 > 3$ to alter the Mg isotopic composition. This may provide a direct observational signature of the operation of the $^{22}{\rm Ne}(\alpha,{\rm n})^{25}$Mg neutron source in stars (e.g. Malaney \& Lambert 1988). Large uncertainties remain, however, in these reaction rates at He-burning temperatures, except for the relatively well-determined \reac{Ne}{20}{\alpha}{\gamma}{Mg}{24} rate. \section{Conclusions} \label{Sect:Conclusions} As an aid to the confrontation between spectroscopic observations and theoretical expectations, the nucleosynthesis associated with the cold CNO, NeNa and MgAl modes of H burning, as well as with He burning, is studied with the help of the recent NACRE compilation of nuclear reaction rates. Special attention is paid to the impact on the derived abundances of the carefully evaluated uncertainties that still affect the rates of many reactions. In order to isolate this nuclear effect in an unambiguous way, a very simple constant temperature and density model is adopted. It is shown that large spreads in the abundance predictions for several nuclides may result not only from a change in temperature, but also from nuclear physics uncertainties. This additional intricacy has to be kept in mind when trying to interpret the observations and when attempting to derive constraints on stellar models from these data. \begin{acknowledgements} This work has been supported in part by the European Commission under the Human Capital and Mobility network contract ERBCHRXCT930339 and the PECO-NIS contract ERBCIPDCT940629. \end{acknowledgements}
\section{Introduction} \label{s1} For scalar field theories in flat space-time, the Osterwalder-Schrader framework provides a valuable link between Euclidean and Minkowskian descriptions of the quantum field. In this paper we will focus only on one aspect of that framework, namely the so-called `reconstruction theorem' \cite{0} which enables one to recover the Hilbert space of quantum states and the Hamiltonian operator, starting from an appropriate measure on the space of Euclidean paths. At least in simple cases, this procedure provides a precise correspondence between the path integral and canonical approaches to quantization. However, since even the basic axioms of the framework are deeply rooted in (Euclidean and) Poincar\'e invariance, a priori it is not obvious that the construction would go through in diffeomorphism invariant theories, such as general relativity. In particular, the idea of a `Wick rotation', implicit in the original framework, has no obvious meaning in this context. The purpose of this paper is to show that, in spite of these difficulties, the construction can be generalized to such theories. While diffeomorphism invariance is our primary concern, we will also address another issue that arises already in Minkowskian field theories. It stems from the fact that the standard Osterwalder-Schrader framework \cite{1} is geared to `kinematically linear' systems ---such as interacting scalar field theories--- where the space of Euclidean paths has a natural vector space structure. More precisely, paths are assumed to belong to the space of Schwartz distributions and this assumption then permeates the entire framework. Although the assumption seems natural at first, in fact it imposes a rather severe limitation on physical theories that one can consider. In particular, in non-Abelian gauge theories, the space of gauge equivalent connections does {\it not} have a natural vector space structure. Therefore, if one wishes to adopt a manifestly gauge invariant approach, the space of histories can not be taken to be one of the standard spaces of distributions \cite{almmt2,almmt1}. Even if one were to use a gauge fixing procedure, because of Gribov ambiguities, one can not arrive at a genuine vector space if the space-time dimensions greater than two. Our extension of the Osterwalder-Schrader reconstruction theorem will incorporate such `kinematically non-linear' theories. Let us now return to our primary motivation. As noted above, since the standard formulation of the reconstruction theorem makes a crucial use of a flat, background metric, it excludes diffeomorphism invariant theories. The most notable examples are gravitational theories such as general relativity and topological ones such as Chern-Simons and BF theories, in which there is no background metric {\it at all}. To incorporate these cases, one has to generalize the very setting that underlies the Osterwalder-Schrader framework. A natural strategy would be to substitute the Poincar\'e group used in the original treatment by the diffeomorphism group. However, one immediately encounters some technical subtleties. In certain cases, such as general relativity on spatially compact manifolds, all diffeomorphisms are analogous to gauge transformations. Hence, while they have a non-trivial action on the space of paths, they have to act trivially on the Hilbert space of physical states. In other cases, such as general relativity in the asymptotically flat context, diffeomorphisms which are asymptotically the identity correspond to gauge while those that preserve the asymptotic structure but act non-trivially on it define genuine symmetries. These symmetries should therefore lead to non-trivial Hamiltonians on the Hilbert space of physical states. The desired extension of the Osterwalder-Schrader framework has to cater to these different situations appropriately. Thus, from a conceptual viewpoint, the extensions contemplated here are very significant. However, it turns out that, \textit{once an appropriate setting is introduced}, the technical steps are actually rather straightforward. With natural substitutions suggested by this generalized setting, one can essentially follow the same steps as in the original reconstruction \cite{1} with minor technical modifications. In particular, it is possible to cater to the various subtleties mentioned above. The plan of this paper is as follows. Section \ref{s2} introduces the new setting and section \ref{s3} contains the main result, the generalized reconstruction theorem. Section \ref{s4} (and the Appendix) discuss examples which illustrate the reconstruction procedure. These examples will in particular suggest a manifestly gauge invariant approach in the non-Abelian context and also bring out different roles that the diffeomorphism group can play and subtleties associated with them. Section \ref{s5} summarizes the main results and briefly discusses their ramifications as well as limitations. \section{The general setting} \label{s2} This section is divided into three parts. In the first, we introduce the basic framework, in the second we discuss some subtleties associated with diffeomorphism invariance and in the third we present the the modified axioms. \subsection{Basic framework} \label{s2.1} Heuristically, our task is to relate the path integral and canonical approaches for a system which may not have a background metric structure. Let us therefore begin with a differential manifold $M$ of dimension D+1 and topology $\Rl\times\sigma$, where $\sigma$ is a D-dimensional smooth manifold of arbitrary but fixed topology. $M$ will serve as the non-dynamical arena for theories of interest. The product topology of $M$ will play an important role in what follows. In particular, it will enable us to generalize the notion of `time-translations' and `time-reflections' which play an important role in the construction. Since our goal is to obtain a Hamiltonian quantum theory, it is not surprising that we have to restrict ourselves to a product topology. Our generalized Osterwalder-Schrader axioms will require the use of several structures associated with $M$. The fact that $M$ is diffeomorphic to $\Rl\times\sigma$ in particular means that it can be foliated by leaves diffeomorphic to $\sigma$. To be precise, consider the set ${\rm Emb}(\sigma,M)$ of all embeddings of $\sigma$ into $M$. A {\it foliation} $E=\{E_t\}_{t\in\Rl}$ is a one-parameter family of elements of ${\rm Emb}(\sigma, M)$, $E_t\in {\rm Emb}(\sigma,M)$, which varies smoothly with $t$ and provides a diffeomorphism between $\Rl \times \sigma$ and $M$. The set of foliations ${\rm Fol}(\sigma,M)$, given by all diffeomorphisms from $\Rl \times \sigma$ to $M$, will be of special interest to us. Notice that the embedded hyper-surfaces $\Sigma_t=E_t(\sigma)$ have not been required to be `space-like, time-like or null.' Indeed, there is no background metric to give meaning to these labels. Each foliation $E\in {\rm Fol}(\sigma,M)$ enables one to generalize the standard notions of time translation and time reflection. To see this, first note that since $E$ is of the form $E: \Rl \times \sigma \rightarrow M; \ (t,x) \mapsto X = E_t(x)$, the inverse map $E^{-1}$ defines functions $t_E(X)$ and $x_E(X)$ from $M$ to $\Rl$ and $\sigma$ respectively. The time translation $\varphi^\Delta$, with $\Delta \in \Rl$, is the diffeomorphism on $M$, $(t_E (X), x_E(X)) \mapsto (t_E(X) + \Delta, x_E(X))$, which is simply a shift of the time coordinate $t_E$ by $t$, holding $x_E(X)$ fixed. Similarly, the time reflection $\theta_E$ is the diffeomorphism of $M$ defined by $(t_E(X),x_E(X))\mapsto (-t_E(X),x_E(X))$. We also consider the positive and negative half spaces $S_E^\pm$, defined by $X \in S^\pm_E$ if and only if $\pm t_E(X) \ge 0$. Although these notions are tied to a specific foliation, our final constructions and results will \textit{not} refer to a preferred foliation. We now turn to the structures associated with the particular quantum field theory under consideration. Let us assume that our theory is associated with a classical Lagrangian density which depends on a collection of basic (bosonic) fields $\phi$ on $M$ and their various partial derivatives. We will not explicitly display discrete indices such as tensorial or representation space indices, so that the symbol $\phi$ may include, in addition to scalar fields, higher spin fields which may possibly take values in a representation of the Lie-algebra of a structure group. The fields $\phi$ belong to a space ${\cal C}$ of {\it classical histories} which is typically a space of smooth (possibly Lie algebra-valued) tensor fields equipped with an appropriate Sobolev norm. In the case of a gauge theory, there will be an appropriate gauge bundle over $M$. We assume that the action of ${\rm Diff}(M)$, the diffeomorphism group of $M$, has a lift to this bundle, from which an action on ${\cal C}$ follows naturally. For notational simplicity, we will denote this action of ${\rm Diff}(M)$ on ${\cal C}$ simply by $\phi\mapsto \varphi \phi$ for any $\varphi \in {\rm Diff}(M)$. Of greater interest than ${\cal C}$ will be the set ${\o{\C}}$ of {\it quantum histories} which is generally an extension of ${\cal C}$. In a kinematically linear field theory, $\o{\C}$ is typically the space of Schwartz distributions \cite{1}. The extension from ${\cal C}$ to $\o{\C}$ is essential because, while ${\cal C}$ is densely embedded in $\o{\C}$ in the natural topology, in physically interesting cases, ${\cal C}$ is generally of measure zero, whence the genuinely distributional paths in $\o{\C}$ are crucial to path integrals. In the more general case now under consideration, we leave the details of the extension unspecified, as they depend on the particulars of the theory being considered, and refer to elements of ${\o{\C}}$ simply as {\it generalized fields}. For example, in a gauge theory these might include generalized connections discussed briefly in Section \ref{s4} (and in detail in \cite{almmt2}). For notational simplicity, the symbol $\phi$ will be used to denote generalized fields (elements of ${\o{\C}}$) as well as smooth fields in ${{\cal C}}$; the context will remove the ambiguity. Consider then a suitable collection of subsets of $\o{\C}$ and denote by ${\cal B}$ the $\sigma$-algebra it generates. This equips $\o{\C}$ with the structure of a measurable space. Let us further consider the set ${\cal F}(\o{\C})$ of measurable functions on this space (that is, functions for which the pre-image of any Lebesgue measurable set is a measurable set). With this background material at hand, we can now introduce a key technical notion, that of a {\it label set} ${{\rm L}}$, which in turn will enable us to define the basic random variables and stochastic process. ${{\rm L}}$ is to be regarded as being `dual' to the space ${\o{\C}}$ of generalized fields. That is, it must be chosen to match the structure of $\overline{{\cal C}}$ so that there is a well-defined `pairing' $P$: \begin{equation} \label{1} P\; :\; {\cal L}\to {\cal F}(\overline{{\cal C}});\; f \mapsto P_f\;. \end{equation} For example, in kinematically linear field theories on $\Rl^{D+1}$, typically $\overline{\cal C}$ is taken to be the space of Schwartz distributions with appropriate tensor and internal indices. Then, ${{\rm L}}$ consists of smooth, rapidly decreasing (test) functions $f$ on $\Rl^{D+1}$, with the pairing $P$ defined by $P_f(\phi)=\exp(i\phi(f))$, where $\phi(f) :=\int d^{D+1}X\,\, \phi(X)f(X)$. In $SU(2)$ gauge theories, a natural candidate for ${{\rm L}}$ is the space of loops on $M$ and the pairing is then defined by $P_f(\phi)= {\rm Tr}\, h_f(\phi)$, the trace of the holonomy $h_f(\phi)$ of the generalized connection $\phi$ around the loop $f$ in a suitable representation of the structure group \cite{almmt1,almmt2}. In general, we will assume that each $f \in {{\rm L}}$ is `associated with' a set supp$(f) \subset M$, which we call the support of $f$. The pairing defines a {\it stochastic process} $f\to P_f(\phi)$, and we refer to $P_f$ as a {\it random variable}. \medskip In the general framework, we will not be concerned with the details of the pairing $P$, but merely ask that it satisfies the following three properties:\\ {\it (A1) The pairing is diffeomorphism covariant, in the sense that there exists a left action of ${\rm Diff}(M)$ on ${\rm L}$, which we denote $f\mapsto (\varphi^{-1}) f$, such that $P_f(\varphi \phi)= P_{(\varphi^{-1}) f}(\phi)$ for any $\phi\in {{\cal C}}$. Furthermore, we require that {\rm supp}$((\varphi^{-1}) f) = \varphi(${\rm supp}$(f))$.}\\ \\ We also introduce a left action of $\varphi$ on random variables: $\varphi(P_f) = P_{\varphi^{-1} f}$. Note that in the familiar case of scalar field theories where the label set is taken to be the set of Schwarz space functions ($f \in {\cal S}$), the action of $\varphi$ on ${\cal L}$ is $(\varphi^{-1}) f = f \circ \varphi^{-1}$. \medskip Let us denote by ${\cal A}$ the set of finite linear combinations ($N<\infty$) of random variables $P_f:$ \begin{equation} \label{2} \psi(\phi):=\sum_{I=1}^N z_I P_{f_I}(\phi) \end{equation} with $z_I\in\Co$ and $f_I\in {{\rm L}}$. The second assumption about the pairing $P$ is :\\ \\ {\it (A2) The vector space $\cal A$ is in fact a $\star$-algebra with unit, whose $\star$ operation is complex conjugation of functions on $\overline {{\cal C}}$. The algebraic operations of ${\cal A}$ must commute with the action of diffeomorphisms in the sense that: }\\ \begin{equation} {\it For} \ a, b \in {\cal A}, \ \varphi(ab) =[\varphi(a)] [\varphi(b)], \ \ {\it and} \ \ [\varphi(a^\star)]= [\varphi(a)]^\star. \end{equation} The first part of this property will allow us to calculate scalar products between elements of ${\cal A}$ with respect to suitable measures on $\overline{{{\cal C}}}$ purely in terms of expectation values of the random variables $P_f$. Note that, in the kinematically linear theories as well as the gauge theories referred to above, this assumption is automatically satisfied. Next, let us consider a $\sigma$-additive probability measure $\mu$ on the measurable space $(\overline{{{\cal C}}},{\cal B})$, thus equipping it with the structure of a measure space $({\o{\C}},{\cal B},\mu)$. This structure naturally gives rise to the so called `history Hilbert space' \begin{equation} \label{3} {\cal H}_{D+1}:=L_2(\overline{{{\cal C}}}, d\mu) \end{equation} of square integrable functions. We denote the inner product between $\psi,\psi'\in {\cal A}$ by \begin{equation} \label{3a} \langle \psi,\psi' \rangle:=\int_{\o{\C}}d\mu \ \overline{\psi(\phi)}\psi'(\phi). \end{equation} Our third requirement on $P$ is that :\\ \\ {\it (A3) The space $\cal A$ is dense in ${\cal H}_{D+1}$ for some measure $\mu$ on $\overline{\cal C}$.}\\ \\ As mentioned above, we are primarily interested in diffeomorphism invariant theories. The pairing allows us to define a representation $\hat{U}(\varphi)$ of ${\rm Diff}(M)$ on the dense subspace ${\cal A}$ of ${\cal H}_{D+1}$: \begin{equation} \label{4} [\hat{U}(\varphi)P_f](\phi):=P_{(\varphi^{-1})f}(\phi)=(\varphi P_f)(\phi). \end{equation} At this point, $\hat{U}(\varphi)$ is a densely defined operator on ${\cal H}_{D+1}$. When the measure $d\mu$ is invariant under diffeomorphisms, we will see that this operator is in fact unitary and extends to all of ${\cal H}_{D+1}$. Finally, in the formulation of the key, `reflection positivity' axiom and in the proof of the reconstruction theorem, we will need certain subsets ${\cal A}_E^\pm$ of $\cal A$. These are defined by restricting the supports of the $f_I$ in (\ref{2}) to be contained in half spaces $S_E^\pm$ on which $\pm t_E \ge 0$. \medskip \subsection{Subtleties} \label{s2.2} We are nearly ready to formulate our extension of the Osterwalder-Schrader axioms. However, our emphasis on diffeomorphism invariant systems will cause a certain change of perspective from the familiar case, e.g. of a kinematically linear field theory in flat space-time. In these simpler theories, the Hamiltonian is an object of primary concern, and its construction is central to the Osterwalder-Schrader reconstruction theorem. Now, we no longer have a background metric and therefore no a priori notion of time translations which Hamiltonians normally generate. An obvious strategy is to treat {\it all} diffeomorphisms as symmetries and seek the corresponding Hamiltonians on the Hilbert space of physical states. However, this turns out not to be the correct procedure because of two subtleties. First, in many diffeomorphism invariant systems, the structure of the classical theory tells us that all diffeomorphisms should be regarded as gauge transformations. This can follow from one of the following three related considerations: i) The initial value formulation could show that the initial data can determine a classical solution {\it only} up to diffeomorphisms; or, ii) For fixed boundary values of fields, the variational principle may provide infinitely many solutions, all related to one another by diffeomorphisms which are identity on the boundary; or, iii) In the Hamiltonian formulation, there may be first class constraints whose Hamiltonian flows correspond to the induced action of ${\rm Diff}(M)$ on the phase space. (Typically, one of these implies the other two.) An example where this is the case is general relativity on a spatially compact manifold. In these cases, one expects quantum states to be diffeomorphism invariant, i.e., the corresponding Hamiltonian operators to vanish identically on the physical Hilbert space. In these theories, then, the reconstruction problem should reduce only to the construction of the Hilbert space of physical states starting from a suitable measure on $\o{\C}$. The second subtlety is that many interesting theories use a background structure. They are therefore {\it not} invariant under the full diffeomorphism group but only under the sub-group which preserves the background structure. An interesting example is provided by the Yang-Mills theory in two dimensional space-times which requires an area form, but not a full metric, for its formulation. The theory is therefore invariant under the group of all area preserving diffeomorphisms, a group which is significantly larger than, say, the Poincar\'e group but smaller than the group of {\it all} diffeomorphisms. (Since the area form is a symplectic form in two dimensions, the group of area preserving diffeomorphisms coincides with the group of symplectomorphisms.) A more common situation is illustrated by general relativity in any space-time dimension, subject to asymptotically flat (or anti-de Sitter) boundary conditions. Here, the background structure consists of a flat (or anti-de Sitter) geometry at infinity. One must therefore restrict oneself to those diffeomorphisms which preserve the specified asymptotic structure. In the general case, we will denote the background structure by $s$ and the sub-group of ${\rm Diff}(M)$ preserving this structure by% \footnote{If there is no background structure, as for example in general relativity or topological field theories on a spatially compact manifold, by ${\rm Diff}(M,s)$ we will mean simply ${\rm Diff}(M)$.} ${\rm Diff}(M, s)$. In the presence of a background structure $s$, our foliations will also be restricted to be compatible with $s$ in the sense that the associated generalized time translations $\varphi_E^t$ and time reflections $\theta_E$ constructed above preserve $s$. Now, given any two foliations $E,\tilde{E}$, there is a unique diffeomorphism $\varphi_{E\tilde{E}}$ on $M$ which maps $E$ to $\tilde{E}$. Note however that even if $E$ and $\tilde{E}$ are compatible with $s$, $\varphi_{E,\tilde{E}}$ need not preserve $s$. This leads us to the following important definitions: \begin{Definition} \label{def1} (a) Two foliations $E$ and $\tilde{E}$ are {\it strongly equivalent} if $\tilde{E}= \varphi_{E \tilde{E}}\circ E$ for some $\varphi_{E \tilde{E}} \in {\rm Diff}(M,s)$.\\ (b) Two foliations $E, \tilde{E}$ will be said to be {\it weakly} equivalent if there exists foliations $E', \tilde{E}'$ which are strongly equivalent to $E, \tilde{E}$ respectively such that the time-reflection maps of $E'$ and $\tilde{E}'$ coincide, i.e. $\theta_{E'}= \theta_{\tilde{E}'}$. \end{Definition} Note that strong equivalence trivially implies weak equivalence but the converse is not true. A simple example which illustrates the difference between these two notions of equivalence is provided by setting $M=\Rl^4$ and choosing the background structure $s$ to be a Minkowskian metric $\eta$. Define $E, \tilde{E}$ as follows: $E: \Rl\times \Rl^3 \rightarrow M; \,(t, x) \mapsto E_t(x) = (t, x)$ and $\tilde{E}: \Rl\times \Rl^3 \rightarrow M; \,(t, x) \mapsto \tilde{E}_t(x) = (bt, x)$, where $b$ is a positive constant. Both of these foliations are compatible with the background structure. However, since the diffeomorphism $\varphi_{E \tilde{E}}$ is not an isometry of $\eta$, the two foliations are {\it not} strongly equivalent. However, they define the same time-reflection map and are therefore weakly equivalent. Strong equivalence of $E$ and $\tilde E$ means that the foliations are in fact related by a symmetry $\varphi_{E \tilde E}$ of the theory and we will see that this symmetry defines a unitary mapping of the physical Hilbert space associated with $E$ to that associated with $\tilde E$ which takes the Hamiltonian generator of $\varphi_E^t$ to that of $\varphi_{\tilde E}^t$. In the case of weak equivalence, the foliations are not related by a symmetry and we should expect no correspondence between the Hamiltonians. The point of this definition, however, is that the construction of the physical Hilbert space itself will depend only on the time inversion map $\theta_E$ induced by the foliation $E$. Thus, when $E$ and $\tilde E$ are weakly equivalent, we will still be able to show that the physical Hilbert spaces are naturally unitarily equivalent, though this equivalence will not of course map the generator of $\varphi^t_E$ to that of $\varphi^t_{\tilde E}$. Finally, it is typical in such theories that certain diffeomorphisms play the role of genuine symmetries while others play the role of gauge. Generally, there is a normal subgroup ${\rm Diff}_G(M,s)$ of ${\rm Diff}(M,s)$ which acts as gauge while the quotient, ${\rm Diff}(M,s)/{\rm Diff}_G(M,s)$, acts as a symmetry group. (In asymptotically flat general relativity, for example, ${\rm Diff}_G(M,s)$ consists of asymptotically trivial diffeomorphisms and the quotient is isomorphic to the Poincar\'e group.) In these contexts, we have a mixed situation: ${\rm Diff}_G(M,s)$ should have a trivial action on the physical Hilbert space, while the action of a symmetry diffeomorphism should be generated by a genuine Hamiltonian as in the original reconstruction theorem. \subsection{Generalized Osterwalder-Schrader Axioms} \label{s2.3} With these subtleties in mind, we can now state our generalization of the Osterwalder-Schrader axioms. Our numbering of the axioms below is chosen to match that of \cite{1}. For flexibility, we wish to allow the possibility that a quantum theory may satisfy only a subset of the following axioms. We will be careful in what follows to explicitly state which axioms are required in order that the various conclusions hold. As in the original construction, the key mathematical object will be a measure $\mu$ on the space of quantum histories. In Minkowskian field theories, $\mu$ can be thought of as a rigorous version of the heuristic measure $\exp -S(\phi)\, {\cal D}\phi$ constructed from the Euclidean action. In the standard Osterwalder-Schrader framework, there are two axioms which are central to the construction of the Hilbert space of states and both are restrictions on the measure $\mu$. The first asks that $\mu$ be Euclidean invariant and the second asks that it satisfy a technical condition called `reflection positivity' formulated in terms of the `time-reflection' operator $\theta$ in the Euclidean space. Given a measure $\mu$ with these properties, one can quotient the space $L^2(\o{\C}, d\mu)$ of square-integrable functions on quantum histories by a certain sub-space, defined by $\theta$, to obtain the Hilbert space of quantum states. Heuristically, the restrictions of quantum histories to the D-dimensional, $t=0$ slice in the Euclidean space define the `quantum configuration space' ${\o{\C}}_{t=0}$ and the quotient enables one to pass from $L^2(\o{\C}, d\mu)$ to the space of square-integrable functions on ${\o{\C}}_{t=0}$. The remaining axioms ensure the existence of a Hamiltonian operator and existence and uniqueness of the vacuum state. In the present context, the time reflection operator $\theta$ is replaced by its generalization $\theta_E$ associated with a foliation and the Poincar\'e group is replaced by ${\rm Diff}(M,s)$. Thus, given any foliation $E$, one can essentially repeat the original construction to obtain a Hilbert space ${\cal H}_D^E$ of physical quantum states. The diffeomorphism invariance of $\mu$ then provides unitary maps relating physical Hilbert spaces constructed from equivalent foliations. \begin{Definition} \label{d1} A quantum theory of fields $\phi\in\o{\C}$ on a space-time $M$ diffeomorphic to $\Rl \times\sigma$ is defined by a probability measure $\mu$ on $\o{\C}$ and a pairing $P$ satisfying (A1), (A2) and (A3) above. The generating functional $\chi$ defined by \begin{equation} \label{5} \chi(f):=<P_f>:=\int_{\o{\C}} d\mu(\phi) P_f(\phi). \end{equation} should satisfy at least the first two of the following axioms: \begin{itemize} \item[(II)] DIFFEOMORPHISM INVARIANCE\\ The measure is diffeomorphism invarian \footnote{ In the case when there are symmetries in the quantum field theory other than diffeomorphism invariance, it would be natural to ask that the measure be invariant under these symmetries as well.} in the sense that, for any $\varphi\in {\rm Diff}(M,s)$, $\chi$ satisfies: \begin{equation} \chi(f)=\chi(\varphi^{-1} f). \end{equation} \item[(III)] REFLECTION POSITIVITY\\ Consider the sesquilinear form on ${\cal A}_E^+$ defined for any $E\in {\rm Fol}(\sigma,(M,s))$ by \begin{equation} \label{6} (\psi,\psi')_E:=\langle \hat{U}(\theta_E)\psi,\psi'\rangle \;. \end{equation} We require $(\psi,\psi)_E\ge 0$ for any $\psi\in {\cal A}_{E}^+$. \item[(GI)]GAUGE INVARIANCE\\ For all $\varphi\in {\rm Diff}_G(M,s)$ and all $\psi\in {\cal A}_{E}^+$ we require \footnote{Again, if there exist gauge symmetries in addition to gauge diffeomorphisms, we ask that these be represented trivially on the physical Hilbert space as well.} \begin{equation} ||\left( \hat{U}(\varphi) - \hat{1}\right) \psi||_E =0,\end{equation} where $|| \ ||_E$ denotes the norm associated with the inner product introduced in axiom III. \item[(I)] CONTINUITY\\ For any $E\in {\rm Fol}(\sigma,(M,s))$ for which the one-parameter group of diffeomorphisms $\varphi_E^t$ does not belong to ${\rm Diff}_G(M,s)$, it acts strongly continuously by operators $\hat{U}(\varphi^t_E)$ on ${\cal H}_{D+1}$. \item[(IV)] CLUSTERING\\ For any $E\in {\rm Fol}(\sigma,(M,s))$ for which $\varphi_E^t$ does not belong to ${\rm Diff}_G(M,s)$, we have that \begin{equation} \label{7} \lim_{t\to\infty} \langle \psi,\hat{U}(\varphi^t_E)\psi'\rangle =\langle \psi,1 \rangle \langle 1,\psi' \rangle \end{equation} for any two $\psi,\psi'\in {\cal A}$. \end{itemize} \end{Definition} Note that if axiom (III), (GI), (I) or (IV) holds for some foliation $E\in {\rm Fol}(\sigma, (M,s))$ then, because of the diffeomorphism invariance (II) of the measure, the axiom also holds for any $\tilde E$ which is strongly equivalent to $E$. Furthermore, if axiom (III) or (GI) holds for some foliation $E\in {\rm Fol}(\sigma, (M,s))$ then it in fact holds for any $\tilde E$ which is {\it weakly} equivalent to $E$. We will conclude this sub-section by comparing these axioms with the standard ones of Osterwalder-Schrader \cite{1}. In axiom (II), we have replaced the Euclidean group in the standard formulation by ${\rm Diff}(M,s)$, and in (I), the time translation group by $\varphi^t_E$. Axiom (I) is usually referred to as the `regularity' axiom and typically phrased as a more technical condition specific to scalar field theories on a flat background \cite{1}. However, as its essential role in that case is to ensure strong continuity of time translations, and as this condition is straightforward to state in the diffeomorphism invariant context, we have chosen to promote this condition itself to the axiom. Finally, note that we have discarded the zeroth `analyticity axiom' of \cite{1} which, roughly speaking, requires the generating functional $\chi$ to be analytic in $f\in {\cal L}$. However, it is not clear that a label space $\cal L$ appropriate to a diffeomorphism invariant or kinematically non-linear field theory should carry an analytic structure. In the standard formulation this axiom allows one to define Schwinger n-point functions in terms of $\chi(f)$. Fortunately, Schwinger functions are not essential to our limited goal of defining the Hilbert space theory. \section{Recovery of the Hilbert Space Theory} \label{s3} The central subject of this section is the following straightforward extension of the classical Osterwalder-Schrader reconstruction theorem \cite{1}. \begin{Theorem} \label{th1} i) For each $E\in {\rm Fol}(\sigma,(M,s))$, axioms (II) and (III) imply the existence of a Hilbert space ${\cal H}^E_D$ of physical states. There is a natural class of unitary equivalences between ${\cal H}^E_D$ and ${\cal H}^{\tilde{E}}_D$ for all $E$ and $\tilde{E}$ in the (weak) equivalence class of $E$. ii) Axioms (I), (II) and (III) imply the existence of self-adjoint operators $[\hat{H}^E]$ on ${\cal H}^E_D$ which generate time translations and which have $[1]_E$ as a vacuum stat \footnote{{\rm That is, a state annihilated by $[\hat{H}^E]$. The brackets on $[\hat{H}^E]$ denote an operator on ${\cal H}^E_D$ as opposed to $L^2({\o{\C}}, d\mu)$ and $[1]_E \in {\cal H}^E_D$ is the equivalence class of elements of ${\cal H}^E_{D+1}$ to which the unit function belongs. See the observation below Eq. (\ref{15}).}}. If the foliations $E$ and $\tilde{E}$ are strongly equivalent, then the operators $[\hat{H}^E]$ and $[\hat{H}^{\tilde{E}}]$ are mapped to each other by a unitary equivalence of the Hilbert spaces ${\cal H}^E_D$ and ${\cal H}^{\tilde{E}}_D$ iii) Axioms (I), (II), (III) and (IV) imply that the vacuum vacuum $[1]_E$ is unique in each ${\cal H}^E_D$. These states are mapped to each other by the unitary equivalence of ii) above. \end{Theorem} We break the proof of this theorem into several lemmas. As noted in the Introduction, the essence of the proof is the same as that in the original Osterwalder-Schrader reconstruction but we present it here for completeness. In the following, $E$ is an arbitrary but fixed foliation compatible with the background structure (if any). \begin{Lemma} \label{la1} By axiom (II), the family of operators, densely defined on ${\cal H}_{D+1}$ for any $\varphi\in {\rm Diff}(M,s)$ by \begin{equation} \label{8} \hat{U}(\varphi)P_f:=P_{(\varphi^{-1}) f} \end{equation} can be extended to a unitary representation of ${\rm Diff}(M,s)$. \end{Lemma} Proof of lemma \ref{la1} :\\ By (A3), $\cal A$ is in fact dense in ${\cal H}_{D+1}$. Since $\hat{U}(\varphi)$ has the inverse $\hat{U}(\varphi^{-1})$ on $\cal A$, it will be sufficient to show that $\hat{U}(\varphi)$ is norm-preserving on $\cal A$ for any $\varphi\in {\rm Diff}(M,s)$ and then to use continuity to uniquely extend it to ${\cal H}_{D+1}$. Recalling condition (A2), for states $a, b \in {\cal A}$, we have \begin{equation} \langle \hat{U}(\varphi) a , \hat{U}(\varphi) b \rangle = \int_{\o{\C}}\, d\mu\, \varphi(a^\star b) \, . \end{equation} But, since the measure is diffeomorphism invariant, this is just the expectation value of $a^\star b$, which is the inner product $\langle a, b \rangle$. $\Box$\\ \begin{Lemma} \label{la2} By axioms (II), (III) the sesquilinear form (\ref{6}) defines a non-negative hermitian form on ${\cal A}^+$. \end{Lemma} Proof of lemma \ref{la2} :\\ The hermiticity follows easily from the fact that $\theta_E\circ\theta_E=\mbox{id}_M$ (so that $[\hat{U}(\theta_E)]^\dagger = \hat{U}(\theta_E)$), unitarity of the $\hat{U}(\varphi)$ as established in lemma \ref{la1}, and the hermiticity of $\langle.,.\rangle$. We have \begin{eqnarray} \label{11} \overline{(\psi,\psi')_E} &=& \overline{\langle \hat{U}(\theta_E)\psi,\psi' \rangle} \ = \ \langle \psi',\hat{U}(\theta_E)\psi \rangle \nonumber\\ &=& \langle \hat{U}(\theta_E)\psi',\psi \rangle \ = \ (\psi',\psi)_E \end{eqnarray} Non-negativity is the content of axiom (III).\\ $\Box$\\ \begin{Lemma} \label{la3} \label{la4} The null space ${\cal N}_E:=\{\psi\in {\cal A}_E^+;\; (\psi,\psi)_E=0\}$ is in fact a linear subspace of ${\cal A}_E^+$ owing to axioms (II), (III). As a result, the form $(.,.)_E$ is well-defined and positive definite on \begin{equation} \label{12} {\cal H}^E_D:=\overline{{\cal A}_E^+/{\cal N}_E}, \end{equation} where the over-line denotes completion with respect to $(.,.)_E$. \end{Lemma} Proof of lemma \ref{la3} :\\ This is a consequence of the Schwarz inequality for positive semi-definite, hermitian, sesquilinear forms.\\ $\Box$\\ \begin{Lemma} \label{la5} The map $[.]_E:\; {\cal A}_E^+ \to {\cal A}_E^+/{\cal N}_E$ is a contraction, that is, $||[\psi]_E||_{{\cal H}^E_D}\le ||\psi||_{{\cal H}_{D+1}}$ owing to axioms (II), (III). \end{Lemma} Proof of lemma \ref{la5} :\\ By the Schwarz-inequality for $\langle.,.\rangle$ and the unitarity of $\hat{U}(\theta_E)$, we have \begin{eqnarray} \label{15} (||[\psi]||_{{\cal H}^E_D})^2 &=& |\langle \hat{U}(\theta_E)\psi,\psi \rangle|^2 \nonumber\\ &\le& ||\hat{U}(\theta_E)\psi||_{{\cal H}_{D+1}} ||\psi||_{{\cal H}_{D+1}} = ||\psi||^2_{{\cal H}_{D+1}}. \end{eqnarray} $\Box$ We will also need the following observation: consider an operator $\hat{A}$ on ${\cal H}_{D+1}$ with dense domain ${\cal D}(\hat{A})$ satisfying the following properties. :\\ Pi) ${\cal D}_E(\hat{A}):=({\cal D}(\hat{A})\cap{\cal A}_E^+)/{\cal N}_E$ is dense in ${\cal H}^E_D$,\\ Pii) $\hat{A}$ maps ${\cal D}(\hat{A})\cap{\cal A}_E^+$ into ${\cal A}_E^+$ and\\ Piii) $\hat{A}$ maps ${\cal D}(\hat{A})\cap{\cal N}_E$ into ${\cal N}_E$.\\ Then the operator $[\hat{A}]:\;{\cal D}_E(\hat{A})\to {\cal H}_D^E$ defined by $[\hat{A}][\psi]:=[\hat{A}\psi]$ is well-defined since $[\hat{A}\psi]=[0]$ for any $\psi\in {\cal N}_E$. We have now prepared all the tools necessary to complete the proof of the theorem. Notice that lemmas \ref{la1} through \ref{la5} have so far used only the axioms (II) and (III). \\ \\ Proof of theorem \ref{th1} :\\ i) For every $E\in {\rm Fol}(\sigma, (M,s))$ we have already constructed ${\cal H}_D^E$. Now, let us suppose two foliations $E$ and $\tilde{E}$ are weakly equivalent. Then, there exist $\varphi_{EE'}, \varphi_{\tilde{E} \tilde{E}'} \in {\rm Diff}(M,s)$ such that $E' = \varphi_{EE'}\circ E$ and $\tilde{E}' = \varphi_{\tilde{E}\tilde{E}'}\circ \tilde{E}$ define the same time-reflection map. Therefore, ${\cal H}_D^{E'} = {\cal H}_D^{\tilde{E}'}$. On the other hand, because of the diffeomorphism $\varphi_{EE'}$, any vector $\psi'\in {\cal A}^+_{E'}$ is of the form $\hat{U}(\varphi_{EE'})\psi$ for some vector $\psi\in {\cal A}_E^+$. Since $\hat{U}(\varphi_{EE'})$ maps ${\cal N}_E$ to ${\cal N}_{E'}$, $\hat{U}(\varphi_{EE'})$ respects the quotient construction and we can define a norm-preserving operator $[\hat{U}_{EE'}]:\;{\cal H}^E_D\to {\cal H}^{E'}_D$ by \begin{equation} \label{16} [\hat{U}_{EE'}][\psi]_E:=[\hat{U}(\varphi_{EE'})\psi]_{E'}. \end{equation} Thus, the Hilbert spaces ${\cal H}^E_D$, ${\cal H}^{E'}_D$ are unitarily equivalent in a natural way. Similarly, there exists a unitary map $[\hat{U}_{\tilde{E} \tilde{E}'}]: {\cal H}_D^{\tilde{E}} \rightarrow {\cal H}_D^{\tilde{E}'}$. Hence, \begin{equation} [\hat{U}_{E \tilde{E}}]:= [\hat{U}_{\tilde{E}\tilde{E}'}]^{-1}\, [\hat{U}_{EE'}] \end{equation} is a natural isomorphism between ${\cal H}_D^E$ and ${\cal H}_D^{\tilde{E}}$. \medskip ii) We will first show the following : \begin{Lemma} \label{la6} For any $E\in {\rm Fol}(\sigma,(M,s))$ and any $t\ge 0$ the operator $\hat{U}(\varphi_E^t)$ satisfies the properties Pi), Pii) and Piii) above and gives rise to a self-adjoint, one-parameter contraction semi-group $[\hat{C}^t_E]$ on ${\cal H}^E_D$. If $\varphi_E^t\not\in\mbox{Diff}_G(M,s)$ then the contraction semi-group is also strongly continuous and its generator $[\hat{H}^E]$ is a self-adjoint, positive semi-definite operator on ${\cal H}^E_D$ with $[1]_E$ a vacuum state for $[\hat{H}^E]$. \end{Lemma} Proof of lemma \ref{la6} :\\ First we show that $\hat{U}(\varphi^t_E)$ has the required properties Pi), Pii), Piii). Consider $f\in {\cal L}$ with supp$(f)\subset S_E^+$. Then supp$((\varphi^t_E)^{-1} f)=\varphi^t_E(\mbox{supp}(f)) \subset \varphi^t_E(S_E^+)=S_E^+$ since for positive $t$ we have a time translation into $S_E^+$ (`the future of $E_0(\sigma)$'). Thus, $\hat{U}(\varphi^t_E)$ maps ${\cal A}_E^+$ into itself. From the Schwarz inequality we infer that it also maps ${\cal N}_E$ into itself and, finally, since ${\cal A}_E^+\subset{\cal A} ={\cal D}(\hat{U}(\varphi^t_E))$ we have ${\cal D}_E(\hat{U}(\varphi^t_E))={\cal A}_E^+/{\cal N}_E$ which is dense in ${\cal H}^E_D$. Thus, the operator \begin{equation} \label{17} [\hat{C}^t_E][\psi]_E:=[\hat{U}(\varphi^t_E)\psi]_E \end{equation} is well defined on a dense domain ${\cal D}_E([\hat{C}^t_E]):={\cal A}_E^+/{\cal N}_E$ of ${\cal H}^E_D$ independent of $t$. That it defines a semi-group follows from the definition (\ref{17}), the fact that the $\varphi^t_E$ form a group under composition and the fact that $\hat{U}$ defines a unitary representation of ${\rm Diff}(M,s)$ on ${\cal H}_{D+1}$. We have \begin{equation} \label{18} [\hat{C}^t_E][\hat{C}^s_E][\psi]_E= [\hat{U}(\varphi^t_E)\hat{U}(\varphi^s_E)\psi]_E =[\hat{U}(\varphi^t_E\circ\varphi^s_E)\psi]_E =[\hat{U}(\varphi^{t+s}_E)\psi]_E =[\hat{C}^{t+s}_E][\psi]_E\;. \end{equation} Next we show that the $[\hat{C}^t_E]$ are Hermitian on ${\cal H}_D^E$. For any $\psi,\psi'\in {\cal D}_E([\hat{C}^t_E])$ we have \begin{eqnarray} \label{19} & & ([\hat{C}^t_E][\psi]_E,[\psi']_E)_E = ([\hat{U}(\varphi^t_E)\psi]_E,[\psi']_E)_E= \langle \hat{U}(\theta_E)\hat{U}(\varphi^t_E)\psi,\psi' \rangle \nonumber\\ & = & \langle \hat{U}(\varphi^t_E)^\dagger\hat{U}(\theta_E)\psi,\psi' \rangle =\langle \hat{U}(\theta_E)\psi,\hat{U}(\varphi^t_E)\psi' \rangle =([\psi]_E,[\hat{C}^t_E][\psi']_E)_E \end{eqnarray} where we have used $\theta_E\circ\varphi^t_E=\varphi^{-t}_E\circ\theta_E$ and the unitarity of the representation of the diffeomorphism group on ${\cal H}_{D+1}$. In particular, we see that ${\cal D}_E([\hat{C}^t_E])$ is contained in ${\cal D}_E([\hat{C}^t_E]^\dagger)$. The contraction property now follows from hermiticity: For $\psi\in {\cal A}_E^+$, we use reflection positivity and find that \begin{eqnarray} \label{20} 0 &\le& ||[\hat{C}^t_E][\psi]_E||_{{\cal H}^E_D}= ([\psi]_E,[\hat{C}^{2t}_E] [\psi]_E)_E^{1/2} \nonumber\\ & \le& ||[\psi]_E||_{{\cal H}^E_D}^{1/2} ||[\hat{C}^{2t}_E][\psi]_E||_{{\cal H}^E_D}^{1/2} \end{eqnarray} from the Schwarz inequality. We see, in particular from the first line of (\ref{20}) that $[\hat{C}^t_E]$ is a positive semi-definite operator. Iterating $n$-times we arrive at \begin{equation} \label{21} ||[\hat{C}^t_E][\psi]_E||_{{\cal H}^E_D} \le ||[\psi]_E||_{{\cal H}^E_D}^{\sum_{k=1}^n(1/2)^k} ||[\hat{C}^{2^nt}_E][\psi]_E||_{{\cal H}^E_D}^{(1/2)^n}\;. \end{equation} Now, using lemma \ref{la5} and again the unitarity of the representation of the diffeomorphism group on ${\cal H}_{D+1}$ we have $||[\hat{C}^{2^nt}_E][\psi]_E||_{{\cal H}^E_D}\le ||\psi||_{{\cal H}_{D+1}}$ and finally since $\sum_{k=1}^n(1/2)^k=1-(1/2)^n$, we find \begin{equation} \label{22} ||[\hat{C}^t_E][\psi]_E||_{{\cal H}^E_D} \le ||[\psi]_E||_{{\cal H}^E_D}^{1-(1/2)^n} ||\psi||_{{\cal H}_{D+1}}^{(1/2)^n}\;. \end{equation} Taking the limit $n\to\infty$ of (\ref{22}) we find the desired result \begin{equation} \label{22a} ||[\hat{C}^t_E][\psi]_E||_{{\cal H}^E_D} \le ||[\psi]_E||_{{\cal H}^E_D}\;. \end{equation} The hermiticity of $\hat{C}^t_E$ together with its boundedness implies that it can be extended to all of ${\cal H}^E_D$ as a self-adjoint, positive semi-definite operator. So far we have made no use of axiom (I). However, it is this axiom that guarantees the existence of a generator for $[\hat C^t_E]$. Note that, for any $\psi\in {\cal H}^E_D$, using the hermiticity of $[\hat{C}^t_E]$, we have \begin{eqnarray} \label{23} 0&\le& ||[\hat{C}^t_E][\psi]_E-[\psi]_E||_{{\cal H}^E_D}^2 \nonumber\\ &=&([\psi]_E,[\hat{C}^{2t}_E][\psi]_E)_E +([\psi]_E,[\psi]_E)_E -2([\psi]_E,[\hat{C}^t_E][\psi]_E)_E \nonumber\\ &=&|\langle \hat{U}(\theta_E)\psi,(\hat{U}(\varphi^{2t}_E)+1- 2\hat{U}(\varphi^t_E))\psi \rangle| \nonumber\\ &\le&| \langle \hat{U}(\theta_E)\psi,(\hat{U}(\varphi^{2t}_E)-1)\psi \rangle| +2| \langle \hat{U}(\theta_E)\psi,(\hat{U}(\varphi^t_E)-1)\psi \rangle| \nonumber\\ &\le& ||\psi||_{{\cal H}_{D+1}} [||(\hat{U}(\varphi^{2t}_E)-1)\psi||_{{\cal H}_{D+1}} +2||\hat{U}(\varphi^t_E)-1)\psi||_{{\cal H}_{D+1}}]\;. \end{eqnarray} Using axiom (I), strong continuity of the one parameter group of unitary operators $\hat{U}(\varphi_E^t)$, we find that the limit of (\ref{23}) vanishes as $t\to 0$, establishing strong continuity of the one-parameter self-adjoint contraction semi-group on ${\cal H}^E_D$. Therefore, using the Hille-Yosida theorem \cite{6} we infer that $[\hat{C}^t_E]=\exp(-t[\hat{H}^E])$ where the generator $[\hat H^E]$ is a positive semi-definite operator on ${\cal H}^E_D$. It must annihilate the state $[1]_E$ as $[\hat{C}^t_E][1]_E=[\hat{U}(\varphi^t_E)1]_E=[1]_E$ for any $t\ge 0$.\\ $\Box$\\ \\ Clearly, if foliations $E,\tilde{E}\in {\rm Fol}(\sigma,(M,s))$ are strongly equivalent, their generators are related by \begin{equation} \label{24} [\hat{H}^{\tilde{E}}]=[\hat{U}(\varphi_{E\tilde{E}})][\hat{H}^E] [\hat{U}(\varphi_{E\tilde{E}}^{-1})]. \end{equation} \\ iii)\\ So far, axiom (IV) has not been invoked. Axiom (IV) tells us that the limit \,\, $\lim_{t\to\infty}\hat{U}(\varphi^t_E)$ becomes the projector $|1 \rangle \langle 1|$ in the weak operator topology. Suppose that there exists a state $\Omega_E$ which is orthogonal to $|1\rangle$ and satisfies $\hat{U}(\varphi^t_E)\Omega_E=\Omega_E$ for any $t\ge 0, E\in {\rm Fol}(\sigma,(M,s))$. We then have \begin{equation} \label{25} ||\Omega_E||^2_{{\cal H}_{D+1}}\, = \, \lim_{t\to\infty} \langle \Omega_E,\hat{U}(\varphi^t_E)\Omega_E \rangle =| \langle 1,\Omega_E \rangle|^2=0. \end{equation} This demonstrates the uniqueness of the vacuum and concludes the proof of the theorem.\\ $\Box$ \\ We will conclude this section with a few remarks. a) The uniqueness result in part ii) of the theorem can be slightly extended. Let $E$ and $\tilde{E}$ be weakly equivalent (rather than strongly, as required in part ii)). Then, there exist $\varphi_{EE'}, \varphi_{\tilde{E}\tilde{E}'} \in {\rm Diff}(M,s)$ such that $\theta_{E'} = \theta_{\tilde{E}'}$. Although, in general the diffeomorphism $\varphi_{EE'}\circ \varphi^{-1}_{\tilde{E}\tilde{E'}}$ will not map the foliation $E$ to $\tilde{E}$, it may map the time translation $\varphi_E^t$ to the time translation $\varphi_{\tilde{E}}^t$. In this case, the unitary map $U_{E,\tilde{E}}$ of part i) of the Theorem will map $[\hat{H}^E]$ to $[\hat{H}^{\tilde{E}}]$ as in (\ref{24}). b) At first it may seem surprising that the uniqueness result for the Hamiltonian is not as strong as in the standard Osterwalder-Schrader construction. However, this is to be expected on general grounds. In the standard construction, the notion of time translation is rigid. In the present context, there is much more freedom. As shown by the example at the end of section \ref{s2.2}, from the viewpoint of the general framework, already in Minkowski space we are led to allow both $\partial / \partial t$ and $b\, \partial/\partial t$ as time evolution vector fields for any positive constant $b$. Clearly, the Hamiltonian operators must also differ by a multiplicative constant in this case. More generally, agreement between Hamiltonians can be expected only if the two generate the same (or equivalent) `translations' in space-time. This is precisely what our uniqueness result guarantees. c) If two foliations $E,\tilde{E}$ are not even weakly equivalent, there may be no natural isomorphism between the physical Hilbert spaces ${\cal H}_D^E$ and ${\cal H}_D^{\tilde{E}}$. In the specific examples we will consider in the next section, all foliations will in fact be weakly equivalent. However, due to the so-called `super-translation freedom' \cite{st}, the situation may well be different in asymptotically flat general relativity in four space-time dimensions. It would be interesting to find explicit examples in which this inequivalence occurs and to understand its physical significance. d) Note that, as in the original Osterwalder-Schrader reconstruction theorem, the Hilbert space theory obtained here is not as complete as one would ideally like it be. In particular, no prescription has been given to construct quantum operators corresponding to the classical (Dirac) observables. \section{Examples} \label{s4} In this section we discuss three examples of measures on spaces of quantum histories. The first example is natural from a mathematical viewpoint but does not obviously come from the path integral formulation of a theory of direct physical interest. (However, we show in the Appendix that, if $D=1$, this measure is naturally associated with a universality class of generally covariant quantum gauge field theories. See \cite{12} for further details.) This measure satisfies some of our axioms. The other two measures satisfy all of our axioms and come from the following systems: Yang-Mills theory in two space-time dimensions and general relativity in three space-time dimensions (or, B-F theory in any space-time dimension). In all three cases, the space $\o{\C}$ of quantum histories is kinematically non-linear and there is no background metric. These examples serve to bring out different aspects of the generalization of the reconstruction theorem. \subsection{The Uniform Measure for Gauge Theories} \label{s4.1} The space of `generalized connections' admits a natural diffeomorphism invariant measure in any space-time dimension, which we will refer to as the {\it uniform measure} and denote by $\mu_{0}$ \cite{7}. It plays a crucial role in the kinematical part of a non-perturbative approach to the quantization of diffeomorphism invariant theories of connections.% \footnote{See \cite{almmt3} for a summary and a discussion of the mathematical details. This approach was motivated in large measure by ideas introduced in \cite{GT} and \cite{CL}.} It has also led to a rich quantum theory of geometry \cite{qg}. As remarked above, if $D\not=1$, it is unlikely that $\mu_0$ would arise as the measure on the space of quantum histories in a theory of direct physical interest. Nonetheless, we discuss it here as a simple example of a measure which satisfies axioms II,III and IV above and to illustrate the construction of the Hilbert space ${\cal H}_D^E$. (For a more complete discussion of this measure, see \cite{almmt3,7}.) Suppose we are interested in a theory of connections based on a compact structure group $K$ on a D+1 dimensional space-time manifold. For simplicity, in this brief account we will set $K=SU(2)$, assume that the principal $K$-bundle over $M$ is trivial, and work with a fixed trivialization. There is no background structure and so ${\rm Diff}(M,s)$ is just ${\rm Diff}(M)$ and any two foliations are strongly equivalent. The gauge subgroup ${\rm Diff}_G(M,s)$ of ${\rm Diff}(M,s)$ depends on the specific theory under consideration. The space $\o{\C}$ of quantum histories is now the moduli space of `generalized connections' defined as follows \cite{8}. Let ${{\cal A}}_{\rm W}$ denote the $C^\star$ algebra generated by Wilson loop functions (i.e. traces of holonomies of smooth connections around closed loops in $M$). $\o{\C}$ is the Gel'fand spectrum of ${{\cal A}}_{\rm W}$. Therefore, it is naturally endowed with the structure of a compact, Hausdorff space and one can show that the moduli space of smooth connections ${\cal C}$ is naturally and densely embedded in $\o{\C}$. The label space $\cal L$ consists of triples $f=(\gamma,\vec{j},\vec{I})$ where $\gamma$ is a graph in $M$, $\vec{j}$ a labeling of its edges with non-trivial irreducible equivalence classes of representations of $K$, and $\vec{I}$ a labeling of its vertices with intertwiners. The stochastic process is defined by $f\mapsto P_f(\phi):=T_{\gamma,\vec{j},\vec{I}}(\phi)$ where the latter is a spin-network function on $\o{\C}$ \cite{5,jb}. (Roughly, each $\phi\in\o{\C}$ assigns to every edge of $\gamma$ a group element, the representations $\vec{j}$ convert these elements into matrices and the function $T_{\gamma, \vec{j}, \vec{I}}$ arises from contractions of indices of these matrices and intertwiners $\vec{I}$. For details, see \cite{5,jb}.) Thanks to this judicious choice of ${\cal L}$, the measure $\mu_0$ can be defined quite simply: \begin{equation} <P_f>=0\,\, {\rm for\,\, all}\,\, f \,\,\, {\rm except}\, f_0=(\emptyset,\vec{0}, \vec{0}),\quad {\rm and} \quad <P_{f_0}>=1. \end{equation} It is easy to see that this measure satisfies axioms (II) and (III) : The only property that one needs to use is that spin-network functions form an orthogonal basis for ${\cal H}_{D+1}$. Given a foliation $E$, the equivalence classes under the quotient by the null vectors are in one to one correspondence with finite linear combinations of spin-network states whose graph lies entirely in the surface $E_0(\sigma)$. This then defines the Hilbert space ${\cal H}^E_D$ which is easily seen to be isomorphic to the Hilbert space defined by the quantum configuration space \cite{8} over $\sigma$ and the corresponding uniform measure $d\mu_{0,\sigma}$. While the uniform measure is associated with certain mathematical models introduced by Husain and Kucha\v{r} \cite{hk}, it does not capture the dynamics of a physical system. Therefore, we have some freedom in the choice of ${\rm Diff}_G(M,s)$. However, if we want to satisfy both the remaining axioms, (GI) and (I), no choice is entirely satisfactory. For example, every diffeomorphism in ${\rm Diff}(M,s)$ has a non-trivial (unitary) action on ${\cal H}_D^E$ so that axiom (GI) is satisfied only if we take the group of gauge diffeomorphisms to be trivial. For this choice of ${\rm Diff}_G(M,s)$, and thus for any other, the measure $d\mu_0$ also satisfies axiom (IV) \cite{9}. However, while ${\rm Diff}(M,s)$ has an unitary action on ${\cal H}_{D+1}$, there is no one-parameter group of diffeomorphisms that acts {\it strongly continuously} on the Hilbert space $L^2(d\mu_0)$ \cite{almmt3}. Thus, axiom (I) is satisfied only for the complementary trivial choice ${\rm Diff}_G(M,s) ={\rm Diff}(M,s)$. Nonetheless, it is true that $P_{f_0}=1$ is the only state invariant under all time translations. \subsection{Two-dimensional Yang-Mills Theory} \label{s4.2} As mentioned in Section \ref{s2.2}, in two space-time dimensions, Yang-Mills action requires only an area 2-form rather than a full space-time metric. Therefore, the theory is invariant under all area preserving diffeomorphisms and thus provides an interesting example for our general framework. Since connected, one-dimensional manifolds without boundary are diffeomorphic either to the circle $S^1$ or to $\Rl$, let us consider Yang-Mills theory with structure group $K = SU(N)$ on space-time manifolds $M = \Rl \times \sigma$ where $\sigma = S^1$ or $\sigma = \Rl$. The background structure $s$ is an area two-form $\omega$ on $M$ and the action reads \begin{equation} \label{B1} S(A) = - \frac{1}{g^2} \int_M {\rm Tr}(F \wedge \star F) = - \frac{1}{g^2} \int_M {dx^0dx^1 \over \omega_{01}} tr(F_{01}^2) \ , \end{equation} where $F$ denotes the curvature two-form of the connection $A$ and $(x^0,x^1) = (t,x)$ are the standard coordinates on $\Rl \times \sigma$. In this case, the group ${\rm Diff}(M,s)$ is the group ${\rm Diff}(M,\omega)$ of area preserving diffeomorphisms. The classical Hamiltonian formulation shows that the gauge transformations of the theory correspond only to local $SU(N)$-rotations. Thus, ${\rm Diff}_{G}(M,\omega)$ contains only the identity diffeomorphism. Finally, the example given in section IIB can be trivially adapted to the case under consideration (simply by replacing $\Rl^3$ by $S^1$ and $\eta$ by $\omega$) to show that there exist compatible foliations which fail to be strongly equivalent. However, it is not difficult to show that all compatible foliations are in fact weakly equivalent. We will now construct a quantum field theory for this system, satisfying all our axioms. For the standard area form, the reconstruction of the Hamiltonian formalism from the Euclidean measure was obtained in \cite{almmt2}. The particular Euclidean measure utilized was the limit as the lattice spacing $a$ goes to zero of the Wilson lattice action for Yang-Mills theory. Let us recall some of the results adapted to the case of a general area form. As label space ${\rm L}$ we use $N$-1-tuples of loops in $M$, $f = (\alpha_1, \cdots , \alpha_{N-1})$. Let $\o{\C}$ be the moduli space of generalized connections as in the previous example and the random process be given by \begin{eqnarray} \label{B2} {\rm L} \ & \rightarrow & \ {\cal F}(\o{\C}) \nonumber \\ f = (\alpha_1, \cdots , \alpha_{N-1}) \ & \mapsto & \ P_{(\alpha_1, \cdots , \alpha_{N-1})}(\phi) := T_{\alpha_1}(\phi) \cdots T_{\alpha_{N-1}}(\phi) \ . \nonumber \end{eqnarray} Here $T_{\alpha}$ denotes the Wilson function, $$ T_{\alpha}(\phi) = {1 \over N} {\rm Tr} (h_\alpha(\phi)) \ , $$ where $h_\alpha$ is the holonomy corresponding to the loop $\alpha$ and the (generalized) connection $\phi$. To begin with, let us consider any compatible foliation $E$. In order to adapt the calculations of \cite{almmt2} we consider a ultraviolet regulator $a$ by taking a (possibly) curved lattice in $M$ made of plaquets diffeomorphic (as manifolds with boundary) to rectangles, with area $a^2$ and such that the time-zero slice $\gamma_E = E_0(\sigma)$ is a union of edges of plaquets. Notice that if $\sigma = S^1$, $\gamma_E$ is a (homotopically non-trivial) loop in $M$. It is easy to verify that the calculations and results of \cite{almmt2} remain essentially the same and that we can take the ultraviolet limit $a \rightarrow 0$ in the expression for the generating functional $\chi(\alpha_1, \cdots , \alpha_{N-1}) = <T_{\alpha_1}(\phi) \cdots T_{\alpha_{N-1}}(\phi)>$. Axioms II and III hold and so we can construct the physical Hilbert spaces. Irrespective of the choice of the compatible foliation $E$, the physical Hilbert space is one-dimensional if $\sigma = \Rl$ and is $L^2(SU(N)/Ad_{SU(N)}, d\tilde \mu_H)$ if $\sigma = S^1$, where $\tilde \mu_H$ is the measure induced on $SU(N)/Ad_{SU(N)}$ by the Haar measure on $SU(N)$. Let us concentrate on the more interesting case of $\sigma = S^1$. From \cite{almmt2} we obtain that the time evolution operator $\hat C_E^t$ is given by \begin{equation} \label{B3} [\hat C_E^t] = e^{{1 \over 2} g^2 {\rm Area}(E,t) \Delta} \ , \end{equation} where $\Delta $ denotes the invariant Laplacian on $SU(N)$ (functions on \break $SU(N)/Ad_{SU(N)}$ can be thought as $Ad_{SU(N)}$-invariant functions on $SU(N)$) and ${\rm Area}(E,t)$ denotes the area inclosed between the loops $\gamma_E = E_0(S^1)$ and $E_t(S^1)$. Thus: \begin{equation} \label{B4} {\rm Area}(E,t) = \int_0^t dt' \int_0^1 dx (E^* \omega)_{01}(x,t') \ . \end{equation} Since $\varphi^u_E = E \circ \varphi^u \circ E^{-1} \in {\rm Diff}(M,\omega) \ , \forall u \in \Rl$ (where $\varphi^u$ denotes the standard time translation on $\Rl \times S^1$) or, equivalently, $\varphi^u \in {\rm Diff}(\Rl \times S^1 , E^* \omega) \ , \ \forall u \in \Rl$, the component $(E^* \omega)_{01}$ does not depend on $t$. Therefore the area in (\ref{B4}) is linear in $t$. Now, we showed in Lemma 5 that $[\hat C_E^t] = exp(- t [\hat H^E])$. Hence, the Hamiltonian can now be read-off as: \begin{equation} \label{B5} [\hat H^E] = - {1 \over 2} g^2 L_E \Delta \ , \end{equation} where \begin{equation} \label{B6} L_E = \int_0^1 dx (E^* \omega)_{01}(x) \ . \end{equation} Notice that if $\tilde E$ is strongly equivalent to $E$, $L_{\tilde E} = L_{E}$ and the two Hamiltonians agree as expected from Theorem 1. What would happen if we use a foliation $\tilde{E}$ which is {\it not} strongly equivalent to $E$? Then, the value of $L_{E}$ (and therefore also the Hamiltonian $[\hat H^{E}]$) will in general change. For example, if we choose, as in section IIB, $\tilde{E}: \Rl\times S^1 \rightarrow M, \, (t,x) \mapsto \tilde{E}_t(x) = (bt,x)$ with $b>0$, then $L_{\tilde E} = b L_{E}$ and therefore $[\hat H^{\tilde E}] = b [\hat H^{E}]$. This is, however, exactly what one would expect since the vector field generating the time translation on $M$ defined by $E$ is $b$ times that defined by $E$. This is a concrete illustration of remark b) at the end of sec \ref{s3}. Finally, in this model, the axiom (GI) holds trivially since ${\rm Diff}_{G}(M, \omega)$ contains only the identity diffeomorphism. Furthermore, all physical states are manifestly $SU(N)$-gauge invariant. The existence of a Hamiltonian operator $[\hat{H}^E]$ implies that axiom (I) holds. Axiom (IV) also holds and the vacuum state is unique. \subsection{2+1 gravity and BF-Theories} \label{s4.3} Fix a 3-manifold $M$ with topology $\Rl \times \sigma$, where $\sigma$ is a compact 2-manifold. In the first order form, the basic fields for general relativity can be taken to be a connection $A$ and a Lie-algebra-valued 1-form $e$. The action is given by \begin{equation} \label{2+1} S(A,e) = \int_{M}\, {\rm Tr}\,\, e\wedge F \end{equation} where the trace is taken in the fundamental representation. If the structure group $K$ is $SO(3)$, we obtain general relativity with signature +,+,+ while if the structure group is $SO(2,1)$, we obtain general relativity with signature -,+,+. The field $e$ can be thought of as a triad, and when $e$ satisfies the equation of motion, the field $A$ is the spin-connection compatible with the triad. The equations of motion on $A$ say that $F$ vanishes. In this case, there is no background structure $s$, ${\rm Diff}_G(M)$ is the connected component of the identity of ${\rm Diff}(M)$, and all foliations are strongly equivalent. The heuristic measure on the space of paths $(A,e)$ is given by `${\cal D}\!A\, {\cal D}\!e\, \exp iS(A,e)$' and if we integrate out the $e$ fields we obtain the measure `$\delta(F)\, {\cal D}\!A$' on the space of connections. This suggests that the rigorous measure should be concentrated on flat connections. It turns out that the moduli space of flat connections is a finite dimensional symplectic manifold% \footnote{There are certain technical subtleties in the $SO(2,1)$ case \cite{10}. In what follows we will assume that a Hausdorff manifold has been obtained by deleting suitable points. The resulting moduli space has disconnected components. By moduli space we will refer either to the `time-like' or `space-like' components.} and therefore has a natural Liouville measure. With this intuitive picture in mind, we will now construct a quantum field theory for this system, satisfying all our axioms. Choose for $\o{\C}$ the moduli space of smooth connections (or a suitable completion thereof. For example, in the $SO(3)$ theory one can use the completion used in example 1.) For ${\cal L}$ we use the space of closed loops $f$ on $M$. The stochastic process is defined by $f\mapsto P_f(\phi) = {\rm Tr}\, h_f(\phi)$ where $h_f(\phi)$ is the holonomy of $\phi \in \o{\C}$ around the closed loop $f$ in $M$ and trace is taken in the fundamental representation. The measure is defined by \begin{equation} <\chi_f>\, =\, <P_f(\phi)> \, =\, \int_{{\cal M}_o} d\mu_L P_f(\phi), \end{equation} where ${\cal M}_o$ is the moduli space of flat connections and $\mu_L$ is the Liouville measure thereon. Note incidentally that, in the resulting history Hilbert space ${\cal H}_{D+1}$, $P_f$ and $P_{f'}$ define the same element if $f$ and $f'$ are homotopic to each other. Hence ${\rm Diff}_G(M)$ is represented by the identity operator on ${\cal H}_{D+1}$. It is straightforward to check that the axioms (II), (III), (GI), (I) and (IV) are all satisfied. (In fact (I) and (IV) hold trivially because ${\rm Diff}_{G}(M,s)$ is so large.) The Hilbert space ${\cal H}_D^E$ is isomorphic to $L^2({\cal M}_o, d\mu_L)$. The Hamiltonian theory can be constructed independently through canonical quantization \cite{11} and yields precisely the same Hilbert space of physical states. Note that the correct correspondence between the path integral and canonical quantization holds for both signatures,\, -,+,+ and +,+,+. However, one has to use measures whose heuristic analogs involve $\exp iS$ in {\it both} cases, so that the signature of the associated metric is not fundamental to determining the heuristic form of the measure. In particular, the Wick rotation has no obvious role in the diffeomorphism invariant context. (For further discussion, see \cite{2}.) This viewpoint is also supported by the fact that 2+1 dimensional general relativity is a special case of B-F theories which can be defined in any dimension and in which there is no natural metric at all; the presence of a metric can thus be regarded as an `accident' of 2+1 dimensions. In these theories, the basic fields are a connection $A$ and a D-1 form $B$ with values in the dual of the Lie algebra. The action has the same form as (\ref{2+1}) with $e$ replaced by $B$. One can repeat essentially the same construction for all of these theories. \section{Discussion} \label{s5} In this paper, we introduced an extension of the Osterwalder-Schrader framework to diffeomorphism invariant theories. The key idea was to generalize the standard setting by dropping all references to the space-time metric. We considered $D+1$ dimensional space-times $M$ with topology $\Rl\times \sigma$, where $\sigma$ is allowed to be an arbitrary, $D$-dimensional manifold. Heuristically, $\Rl$ serves as a generalized `time direction'. More precisely, using foliations $E$ of $M$, with leaves transverse to the $\Rl$-direction, we were able to extend the standard notions of time translation and time reflection without any mention of a space-time metric. This in turn enabled us to generalize the Osterwalder-Schrader axioms and construct a Hamiltonian quantum theory starting from a path integral. While $M$ is required to have a product topology, given our goal of constructing a Hamiltonian framework, this restriction is unavoidable. As in the original Osterwalder-Schrader framework, the key mathematical object in the path integral formulation is the measure $\mu$ on the space $\o{\C}$ of quantum histories and the axioms are restrictions on permissible $\mu$. In the construction of a bridge from the path integral to the Hilbert space theory, two of these axioms play a central role: reflection positivity (axiom III, unchanged from the original Osterwalder-Schrader treatment) and diffeomorphism invariance (axiom II, which replaces the Euclidean invariance of the standard treatment). Given a foliation $E$ of $M$ and a measure $\mu$ satisfying reflection positivity, one can construct the Hilbert space ${\cal H}^E_D$ of quantum states. The diffeomorphism invariance of $\mu$ then ensures that the Hilbert space is essentially insensitive to the choice of the foliation $E$. The remaining axioms ensure the existence of the Hamiltonian operators generating (generalized) time-translations which are true (i.e. non-gauge) symmetries of the theory and the existence and uniqueness of a vacuum state. Perhaps the most striking feature of the present framework is its generality. We did not have to restrict ourselves to specific space-time manifolds and the Lagrangian --- indeed even the matter content--- of the theory was left arbitrary. In particular, our generalized setting allows theories of interacting gauge and tensor fields with arbitrary index structure, general relativity, higher derivative gravity theories, etc. However, this generality comes at a price. As with the original Osterwalder-Schrader reconstruction theorem, the results of this paper only tell us how to obtain a Hilbert space theory from a given measure satisfying certain axioms. It does \textit{not} tell us how to construct this measure from a given classical theory. For familiar field theories (without diffeomorphism invariance), $\exp -\,S_E$, with $S_E$, the Euclidean action, generally provides a heuristic guide in the construction of this measure. We saw in Sections \ref{s4.2} and \ref{s4.3} that, in the absence of a space-time metric, the distinction between the usual Euclidean and Lorentzian prescriptions become blurred. In some cases, the heuristic guide is again provided by $\exp -S$ while in other cases it is provided by $\exp iS$. Thus, the construction of the measure now acquires a new subtle dimension. Furthermore, because our setting is much more general than the original one, even the `kinematical structure' ---the spaces $\o{\C}$ of quantum histories, the label set ${\cal L}$ and pairings $P$ of Section \ref{s2}--- can vary from one theory to another and have to be constructed case by case. However, for diffeomorphism theories of connections, including general relativity in three space-time dimensions, we were able to provide natural candidates for these structures and find the appropriate measures. In \cite{2}, we will extend these considerations to more general contexts, albeit at a more heuristic level. In particular, starting from the classical Hamiltonian framework, we will discuss how one can construct heuristic measures. We will find some subtle but important differences from the familiar cases. \bigskip\bigskip {\bf Acknowledgments:} We thank Jerzy Lewandowski for stimulating discussions, Olaf Dreyer for comments on an early draft and Jorge Pullin for numerous suggestions which significantly improved the final presentation. This research project was supported in part by the National Science Foundation under grant PHY95-14240 to Penn State, PHY97-22362 to Syracuse University and PHY94-07194 to ITP, Santa Barbara, by the Eberly research funds of Penn State, by research funds from Syracuse University, by CENTRA/IST, by projects PRAXIS/2/2.1/FIS/286/94 and CERN/P/FIS/1203/98. \begin{appendix} \section{A Universality Class of Generally Covariant Quantum Gauge Field Theories in Two Spacetime Dimensions} \label{sa} The purpose of this appendix is to show that, if $D=2$, the uniform measure $\mu_0$ on the space $\overline{\cal C}$ of generalized connections considered in section \ref{s4.1} naturally arises in the path integral formulation of a class of diffeomorphism invariant gauge theories. For generalizations and further discussion, see \cite{12}. Let $M$ be a two-dimensional manifold with topology $\Rl^2$ or $S^1\times \Rl$, and let $G$ be a compact, connected, semi-simple gauge group. We will denote by $A$ the pull-back by local sections of a connection on a principal $G$-bundle over $M$ and by $F$ its curvature. We will use $a,b,c,..=1,2$ as the tensorial indices and $i,j,k,..=1,..,\dim(G)$ as the Lie algebra indices. Choose a basis $\tau_i$ of the Lie algebra $Lie(G)$ of $G$ normalized such that tr$(\tau_i\tau_j)=-N\delta_{ij}$ and structure constants are defined by $[\tau_i,\tau_j]=2f_{ij}\;^k \tau_k$. Finally, let $\epsilon^{ab}$ be the metric independent totally skew tensor density of weight one and use it to represent the curvature as a Lie-algebra valued scalar density $F^i:=\frac{1}{2}\epsilon^{ab} F_{ab}^i$. Let us consider the following action \begin{equation} \label{a.1} S= \int_M \,d^2x \,\left[F^i F^i\right]^{1\over 2} \equiv \int_M \, d^2x \left[P_2(F)\right]^{1\over 2} \end{equation} where $F^iF^i=k_{ij}F^iF^j$ is the norm of $F^i$ with respect to the Cartan-Killing metric. Since the quantity under the square-root is a scalar density of weight two, the action (\ref{a.1}) is diffeomorphism invariant. This is perhaps the simplest of such actions for $G$-connections. Other, more general diffeomorphism invariant actions can be constructed by taking n-th roots of suitable n-nomials in $F^i$ and their covariant derivatives. Assuming appropriate boundary conditions, the equations of motion that follow from (\ref{a.1}) are \begin{equation} \label{a.2} D_a\frac{F^i}{\sqrt{F^j F^j}}=0 \end{equation} These equations are consistent since the integrability condition $\epsilon^{ab} D_a D_b (F^i/\sqrt{F^j F^j})= f^i\;_{jk} F^j F^k/\sqrt{F^l F^l}=0$ is identically satisfied. The general solution is $F^i=s b^i$ where $s$ is an arbitrary, nowhere negative, scalar density of weight one and $b^i$ a covariantly constant vector of unit norm. In the special case $G=U(1)$, we have: $b^i(x)=b^1(x)=\pm 1$ must be constant, say $b^1(x)=1$ and thus $F(x)=\epsilon^{ab} (\partial_a A_b)(x) \ge 0$ is the general solution. The case $G=U(1)$ also gives a simple class of solutions for general $G$ : Suppose we choose the connection to be of the form $A_a^i=a_a t^i$ where $t^i$ is a constant unit norm vector. Then $F^i=F t^i$ where $F=\epsilon^{ab} \partial_a a_b$ and $D_a F^i/{\sqrt{F^j F^j}}=t^i \partial_a F/|F|$ and this reduces the problem to $G=U(1)$. Thus, the space of solutions to the field equations is infinite dimensional. We now wish to derive a path integral for this theory, i.e., construct a continuum measure along the constructive approach of \cite{almmt2}, using the action (\ref{a.1}) as the classical input. Let us focus on the non-trivial case when $M$ is topologically $\Rl\times S^1$. Let us introduce a system of coordinates $(x,t)$ where $x\in[-a,a]$ denotes the compact direction and $a$ is an arbitrary parameter of dimension of length. Next we introduce a foliation of the cylinder by circles of constant `time' $t$ coordinate. We introduce a dimensionless UV cut-off $\epsilon$ and an IR cut-off $T$ of dimension of length. Consider $1+2N,N=T/(\epsilon a)$ circles at values of $t$ given by $t/a=l\epsilon,\;l=-N,-N+1,..,N$ and similarly $2N'+1,N'=1/\epsilon$ coordinate lines at constant values of $x$ given by $x/a=k\epsilon,\;k=-N',-N'+1,.., N'$ with $x/a=\pm 1$ identified. Thus we have covered a portion $M_T$ of the cylinder $M$, corresponding to $(x,t)\in [-a,a]\times [-T,T]$ by a lattice $\gamma_{\epsilon,T}$ of cubic topology and can consider the usual plaquette loops $\Box$ of that lattice based at the point $p$ with coordinates $x=t=0$, say. We can label plaquettes by two integers $(k,l)$ in the obvious way and have \begin{equation} \label{a.3} \Box_{k,l}=\rho_{k,l}\circ e_{k,l}\circ f_{k+1,l}\circ e_{k,l+1}^{-1} \circ f_{k,l}^{-1}\circ\rho_{k,l}^{-1} \end{equation} where $e_{k,l},\;f_{k,l}$ are respectively edges of the lattice at constant values of $t$ and $x$ respectively and $\rho_{k,l}$ is an arbitrary but fixed lattice path between the base point and the corner of the box corresponding to lowest values of $x,t$ respectively (modulo the identification of $x/a=\pm 1$). If we now parameterize edges as the image of the interval $[0, a]$ then it is not difficult to see that for a classical connection $A$ the holonomy for a plaquette is to leading order in $\epsilon$ given by \begin{equation} \label{a.4} h_{\Box_{k,l}}=1+a^2\epsilon^2 \epsilon_{ab}\dot{e}_{k,l}^a(0) \dot{f}_{k,l}^b(0) F^j(x=k\epsilon a,t=l\epsilon a)\tau_j/2 \end{equation} Then, \begin{equation} \label{a.5} S_{\epsilon,T}:= \sum_{k,l} \left[ P_2(-\frac{2}{N} \mbox{tr}(\tau_i(h_{\Box_{k,l}}-1)))\right]^{1\over 2}=: \sum_{k,l} s(h_{\Box_{k,l}}) \end{equation} can be taken to be a Wilson-like action that approximates (\ref{a.1}) in the sense that \begin{equation} \label{a.6} \lim_{\epsilon\to 0} S_{T,\epsilon}= \int_{M_T}\, d^2x\, \left[{P_2}(F)\right]^{1\over 2} \end{equation} The elementary but important observation is that function $s(h)$ defined in (\ref{a.5}) is the same for all plaquette loops. Let us choose as our random variables the `loop network functions' $T_{\vec{\alpha},\vec{\pi},c}$ \cite{13}. These are similar to the $T_{\gamma, \vec{j}. \vec{I}}$ considered in section \ref{s4.1}), except that: i)now the graphs $\alpha$ are replaced by a finite collection, say $L$, of mutually non-overlapping loops $\vec{\alpha}$ based at $p$ (possibly including the homotopically non-trivial one that wraps once around the cylinder at $t=0$); ii) $\vec{\pi}$ now denotes a collection of $L$ equivalence classes of non-trivial irreducible representations of $G$ subject to the constraint that if loops $\alpha_{I_1},..,\alpha_{I_r},r\le L,I_1<..<I_r$ share a segment, then the tensor product $\pi_{I_1}\otimes...\otimes \pi_{I_r}$ does not contain a trivial representation; and, iii) $c$ is an intertwiner between the trivial representation and $\pi_1\otimes..\otimes\pi_L$. The function $T_{\vec{\alpha},\vec{\pi},c}(A)$ depends on $A$ through the holonomies $h_{\alpha_I}(A),\;I=1,..,L$ only. Given any measure on the moduli space $\overline{\cal C}$ of generalized connections, one can compute the the expectation values of these loop-network functions. These provide us with the characteristic function of the underlying measure \cite{7,almmt2}. Using this machinery, we can write down the regularized measure on $\overline{\cal C}$ by specifying its characteristic function: \begin{equation} \label{a.7} <T_{\vec{\alpha},\vec{\pi},c}>_{T,\epsilon} :=\frac{1}{Z_{\epsilon,T}} \prod_{k,l} ([\int_G d\mu_H(h_{e_{k,l}})]\;[\int_G d\mu_H(h_{e_{k,l}})]) e^{-S_{\epsilon,T}} T_{\vec{\alpha},\vec{\pi},c}\, , \end{equation} where, \begin{equation} Z_{\epsilon,T}:= \prod_{k,l} ([\int_G d\mu_H(h_{e_{k,l}})]\;[\int_G d\mu_H(h_{e_{k,l}})]) e^{-S_{\epsilon,T}}\, . \end{equation} Here, of course, all the loops in question live on our lattice. In order to explicitly compute the expectation value (\ref{a.7}) in the limit $\epsilon\to 0$ and $T\to\infty$ we can essentially follow \cite{almmt2}. This is due to a peculiarity of $D=1$ and the planar or cylindrical topology of $M$ namely, that the plaquette loops are holonomically independent. Let then, at fixed $\epsilon,T$, $|\alpha|$ be the number of plaquettes contained in the surface bounded by $\alpha$. Then, repeating literally all the calculations performed in \cite{almmt2} we find \begin{eqnarray} \label{a.8} <T_{\vec{\alpha},\vec{\pi},c}>_{T,\epsilon} &=& \,\,T_{\vec{\alpha},\vec{\pi},c}(A=0) \prod_{I=1}^L (\frac{J_{\pi_I}}{J_{\pi_0}})^{|\alpha_I|}\nonumber\\ J_\pi &=& \frac{1}{\dim(\pi)}\int_G d\mu_H(h) \chi_\pi(h) e^{-s(h)} \end{eqnarray} Here $\chi_\pi$ denotes the character of $\pi$, $\chi_\pi(1)=\dim(\pi)$ and $\pi_0$ denotes the equivalence class of the trivial representation. If one of the loops, say $\alpha_I$, is homotopically non-trivial then $(\frac{J_{\pi_I}}{J_{\pi_0}})^{|\alpha_I|}$ has to be replaced by $\delta_{\pi_I,\pi_0}$. The UV and IR cut-off can now be trivially removed from (\ref{a.8}) : Due to the holonomic independence of the plaquette loops the quantity (\ref{a.8}) is already independent of $T$. Due to the diffeomorphism invariance of the original action, the quantity (\ref{a.8}) depends on $\epsilon$ only through the numbers $|\alpha_I|$ which just counts the number of plaquette loops that one uses in order to approximate the loop $\alpha_I$. In contrast to the situation with two-dimensional Yang-Mills theory \cite{almmt2}, the numbers $J_\pi$ are already independent of $\epsilon$. The reason for this is, of course, the background independence of the original action (\ref{a.1}). By contrast, as we saw in section \ref{s4.2}, the Yang-Mills action requires a background area-element. At first the difference seems small. However, it leads one in Yang-Mills theory to the Wilson action $\frac{1}{\epsilon^2} \sum_\Box [s(h_\Box)]^2$ ---rather than $\sum_\Box s(h_\Box)$--- which makes $J_\pi$ $\epsilon$-dependent in just the right way to produce the area law \cite{almmt2} as $\epsilon\to 0$. In the present case there is no background structure and therefore we cannot have an area law; there is no area 2-form to measure the area with! Instead we have the following. Since the definition of $J_\pi$ implies $|J_\pi|<J_{\pi_0}$ for $\pi\not=\pi_0$ and since in the limit $\epsilon\to 0$ the numbers $|\alpha_I|$ diverge, it follows that $\lim_{\epsilon\to 0} (J_{\pi_I}/J_{\pi_0})^{|\alpha_I|}=\delta_{\pi,\pi_0}$. Consequently, we have: \begin{equation} \label{a.9} \lim_{T\to\infty} \lim_{\epsilon\to 0} <T_{\vec{\alpha},\vec{\pi},c}>_{T,\epsilon} =\left\{ \begin{array}{cc} 1 & \mbox{ if } \vec{\alpha}=\{p\},\vec{\pi}=c=\pi_0 \\ 0 & \mbox{ otherwise} \end{array} \right. \end{equation} in other words, we arrive at the characteristic functional of the uniform measure in $D=1$ discussed in section \ref{s4.1}. Note that the limit (\ref{a.9}) is {\it completely insensitive to the choice of the regularizing lattice}. Hence, the result is independent of the regulator. Finally, we note that in place of (\ref{a.1}) we could have considered the most general diffeomorphism and gauge invariant action $\tilde{S}(A)$ which depends on the field strengths $F^i$ but not on their derivatives. For this, we can begin with globally defined (i.e. gauge invariant) monomials $F^n:=F^{i_1}..F^{i_n}\mbox{tr}(\tau_{i_1}..\tau_{j_1})$. (If the rank of $G$ is $R$, only the first $R-1$ of these will be independent, others being polynomials in them.) Let $P_n$ be an arbitrary, gauge invariant, positive semi-definite, homogeneous polynomial of degree $n$ in the $F^j$ and set \begin{equation} \label{a.10} \tilde{S}(A)=\sum_{n=1}^\infty a_n \int_M d^2x \sqrt[n]{P_n[F]} \end{equation} where $a_i$ are non-negative constants all but a finite number of which are zero. This action is diffeomorphism invariant, again because the integrand is a scalar density of weight one. We could have carried out the above procedure for any of these theories. Irrespective of the action in this class, the final characteristic function would have been again (\ref{a.9}). Thus, all these theories lie in the same universality class; their renormalization group flows all reach the same UV fixed point and the final quantum theory is dictated by the uniform measure. This issue, the Hamiltonian formulation, and the canonical quantization of these theories will be discussed in \cite{12}. \end{appendix}
\section*{Introduction} Fulton asked how many solutions to a problem of enumerative geometry can be real, when that problem is one of counting geometric figures of some kind having specified position with respect to some general fixed figures~\cite{Fu_84}. For the problem of plane conics tangent to five general conics, the (surprising) answer is that all 3264 may be real~\cite{RTV}. Similarly, given any problem of enumerating $p$-planes incident on some general fixed subspaces, there are real fixed subspaces such that each of the (finitely many) incident $p$-planes are real~\cite{So99}. We show that the problem of enumerating parameterized rational curves in a Grassmannian satisfying simple (codimension 1) conditions may have all of its solutions be real. This problem of enumerating rational curves on a Grassmannian arose in at least two distinct areas of mathematics. The number of such curves was predicted by the formula of Vafa and Intriligator~\cite{Vafa,Intriligator} from mathematical physics. It is also the number of complex dynamic compensators which stabilize a particular linear system, and the enumeration was solved in this context~\cite{RRW98,RRW96}. The question of real solutions also arises in systems theory~\cite{Byrnes}. Our proof, while exploiting techniques from systems theory, has no direct implications for the problem of real dynamic output compensation. \section{Statement of results} We work with complex algebraic varieties and ask when {\it a priori}$\,$ complex solutions to an enumerative problem are real. Fix integers $m,p>1$ and $q\geq 0$. Set $n:=m+p$. Let ${\bf G}$ be the Grassmannian of $p$-planes in ${\mathbb C}^n$. The space ${\mathcal M}_q$ of maps $M:{\mathbb P}^1\rightarrow{\bf G}$ of degree $q$ has dimension $N:=pm+qn$~\cite{Clark,Stromme}. If $L$ is an $m$-plane and $s\in {\mathbb P}^1$, then the collection of all maps $M$ satisfying $M(s)\cap L\neq \{0\}$ is an irreducible subvariety of codimension 1. We study the following enumerative problem: \begin{equation} \label{eq:enumerative} \begin{minipage}{5.2in} Given general points $s_1,\ldots,s_N$ in ${\mathbb P}^1$ and general $m$-planes $L_1,\ldots,L_N$ in ${\mathbb C}^n$, how many maps $M\in {\mathcal M}_q$ satisfy $M(s_i)\cap L_i\neq\{0\}$ for $i=1,\ldots,N$? \end{minipage} \end{equation} Rosenthal~\cite{Rosen94} interpreted the solutions as a linear section of a projective embedding of ${\mathcal M}_q$, and Ravi, Rosenthal, and Wang~\cite{RRW98,RRW96} show that the degree of its closure ${\mathcal K}_q$ in this embedding is \begin{equation}\label{dqmp} \delta\quad:=\quad(-1)^{q(m+1)}N! \sum_{\nu_1+\cdots+\nu_p=q} \frac{\prod_{i<j}(j-i+n(\nu_j-\nu_i))} {\prod_{j=1}^p(m+j+n\nu_j-1)!}\ . \end{equation} Thus, if there are finitely many solutions, then their number (counted with multiplicity) is at most $\delta$. The difference between $\delta$ and the number of solutions counts points common to both the linear section and the boundary ${\mathcal K}_q-{\mathcal M}_q$ of ${\mathcal K}_q$. Since ${\bf G}$ is a homogeneous space, an application of Kleiman's Theorem~\cite{Kleiman} shows there are finitely many solutions and no multiplicities. Bertram~\cite{Bertram} uses explicit methods (a moving lemma) to to show there are finitely many solutions and also no points in the boundary of ${\mathcal Q}_q$, and hence none in the boundary of ${\mathcal K}_q$. He also computes the small quantum cohomology ring of ${\bf G}$, which gives algorithms for computing $\delta$ and other intersection numbers involving rational curves on a Grassmannian. When the $s_i$ and $L_i$ are real, not all of these solutions are defined over the real numbers. We show there are real $s_i$ and $L_i$ for which each of the $\delta$ maps are real. \begin{thm}\label{thm:real} There exist real $m$-planes $L_1,\ldots,L_N$ in ${\mathbb R}^n$ and points $s_1,\ldots,s_N\in{\mathbb P}^1_{\mathbb R}$ so that there are exactly $\delta$ maps $M:{\mathbb P}^1\rightarrow{\bf G}$ of degree $q$ which satisfy $M(s_i)\cap L_i \neq \{0\}$ for each $i=1,\ldots,N$, and each of these are real. \end{thm} Our proof is elementary in that it argues from the equations for the locus of maps $M$ which satisfy $M(s)\cap L\neq \{0\}$. A consequence is that we obtain fairly explicit choices of $s_i$ and $L_i$ which give only real maps, which we discuss in Section 4. Also, our proof uses neither Kleiman's Theorem nor Bertram's moving lemma, and thus it provides a new and elementary proof that there are $\delta$ solutions to the enumerative problem~(\ref{eq:enumerative}). \section{The quantum Grassmannian} The space ${\mathcal M}_q$ of maps ${\mathbb P}^1\rightarrow{\bf G}$ of degree $q$ is a smooth quasi-projective algebraic variety. A smooth compactification is provided by a quot scheme ${\mathcal Q}_q$~\cite{Stromme}. By definition, there is a universal exact sequence $$ 0\ \rightarrow\ {\mathcal S}\ \rightarrow\ {\mathbb C}^n\otimes{\mathcal O}\ \rightarrow\ {\mathcal T}\ \rightarrow 0 $$ of sheaves on ${\mathbb P}^1\times{\mathcal Q}_q$ where ${\mathcal S}$ is a vector bundle of degree $-q$ and rank $p$. Twisting the determinant of ${\mathcal S}$ by ${\mathcal O}_{{\mathbb P}^1}(q)$ and pushing forward to ${\mathcal Q}_q$ induces a Pl\"ucker map $$ {\mathcal Q}_q\ \rightarrow\ {\textstyle {\mathbb P}\left( \bigwedge^p{\mathbb C}^n\otimes H^0({\mathcal O}_{{\mathbb P}^1}(q))^*\right)} $$ which is the analog of the Pl\"ucker embedding of ${\bf G}$. The Pl\"ucker map is an embedding of ${\mathcal M}_q$, and so its image ${\mathcal K}_q$ provides a different compactification of ${\mathcal M}_q$. We call ${\mathcal K}_q$ the quantum Grassmannian. (In~\cite{BDW}, this space is called the Uhlenbeck compactification). Our proof of Theorem~\ref{thm:real} exploits some of its structures that were elucidated in work in systems theory. The Pl\"ucker map fails to be injective on the boundary ${\mathcal Q}_q-{\mathcal M}_q$ of ${\mathcal Q}_q$. Indeed, Bertram~\cite{Bertram} constructs a ${\mathbb P}^{p-1}$ bundle over ${\mathbb P}^1\times{\mathcal Q}_{q-1}$ that maps onto the boundary, with its restriction over ${\mathbb P}^1\times{\mathcal M}_{q-1}$ an embedding. On this projective bundle, the Pl\"ucker map factors through the base ${\mathbb P}^1\times{\mathcal Q}_{q-1}$ and the image of a point in the base is $s\cdot S$, where $s$ is the section of ${\mathcal O}_{{\mathbb P}^1}(1)$ vanishing at $s\in{\mathbb P}^1$ and $S$ is the image of a point in ${\mathcal Q}_{q-1}$ under its Pl\"ucker map. This identifies the image of the exceptional locus of the Pl\"ucker map with the image of ${\mathbb P}^1\times{\mathcal K}_{q-1}$ in ${\mathcal K}_q$ under a map $\pi$ (given below). More concretely, a point in ${\mathcal Q}_q$ may be (non-uniquely) represented by a $p\times n$-matrix $M$ of forms in $s,t$, with homogeneous rows and whose maximal minors have degree $q$~\cite{RR94}. The image of such a point under the Pl\"ucker map is the collection of maximal minors of $M$. The maps in ${\mathcal M}_q$ are represented by matrices whose maximal minors have no common factors: Given such a matrix $M$, the association $$ {\mathbb P}^1\ni(s,t)\ \longmapsto\ \mbox{row space }M(s,t) $$ defines a map of degree $q$. The collection $\binom{[n]}{p}$ of $p$-subsets of $\{1,\ldots,n\}$ index the maximal minors of $M$. For $\alpha\in\binom{[n]}{p}$ and $0\leq a\leq q$, the coefficients $z_{\alpha^{(a)}}$ of $s^at^{q-a}$ in the $\alpha$th maximal minor of $M$ provide Pl\"ucker coordinates for maps in ${\mathcal M}_q$, and for the space ${\mathbb P}\left(\bigwedge^p{\mathbb C}^n\otimes H^0({\mathcal O}_{{\mathbb P}^1}(q))^*\right)$. Let ${\mathcal C}_q:= \{\alpha^{(a)}\mid \alpha\in\binom{[n]}{p}, 0\leq a\leq q\}$ be the indices of these Pl\"ucker coordinates. Then the image of the exceptional locus in ${\mathcal K}_q$ is the image of the (birational) map $\pi: {\mathbb P}^1\times{\mathcal K}_{q-1}\rightarrow{\mathcal K}_q$ defined by \begin{equation}\label{eq:pi} \pi\ :\ \left([A,B],(x_{\beta^{(b)}}\mid \beta^{(b)}\in{\mathcal C}_{q-1})\right) \ \longmapsto\ (Ax_{\alpha^{(a)}}-Bx_{\alpha^{(a-1)}}\mid \alpha^{(a)}\in{\mathcal C}_q)\,. \end{equation} \rule{0pt}{15pt} The relevance of the quantum Grassmannian ${\mathcal K}_q$ to the enumerative problem~(\ref{eq:enumerative}) is seen by considering the condition for a map $M\in{\mathcal M}_q$ to satisfy $M(s,t)\cap L\neq\{0\}$ where $L$ is an $m$-plane in ${\mathbb C}^n$ and $(s,t)\in{\mathbb P}^1$. If we represent $L$ as the row space of a $m\times n$ matrix, also written $L$, then this condition is $$ 0\ =\ \det\left[\begin{array}{c}L\\M(s,t)\end{array}\right]\ =\ \sum_{\alpha\in\binom{[n]}{p}} f_\alpha(s,t)\, l_\alpha\,, $$ the second expression given by Laplace expansion of the determinant along the rows of $M$. Here, $l_\alpha$ is the appropriately signed maximal minor of $L$. If we expand the forms $f_\alpha(s,t)$ in this last expression, we obtain $$ \sum_{\alpha^{(a)}\in{\mathcal C}_q} z_{\alpha^{(a)}} s^a t^{q-a}l_\alpha\ =\ 0\,, $$ a linear equation in the Pl\"ucker coordinates of $M$. Thus the solutions $M\in{\mathcal M}_q$ to the enumerative problem~(\ref{eq:enumerative}) are a linear section of ${\mathcal M}_q$ in its Pl\"ucker embedding, and so the degree $\delta$ of ${\mathcal K}_q$ provides an upper bound on the number of solutions. The set ${\mathcal C}_q$ of Pl\"ucker coordinates has a natural partial order $$ \alpha^{(a)}\ \geq\ \beta^{(b)} \quad\Longleftrightarrow\quad \begin{array}{c} a\geq b, \mbox{ and if } a-b<p, \mbox{ then }\\ \alpha_{a-b+1}\geq \beta_1, \ldots,\alpha_p\geq \beta_{p+1-b+a} \end{array}\ . $$ The poset ${\mathcal C}_q$ is graded with the rank, $|\alpha^{(a)}|$, of $\alpha^{(a)}$ equal to $an + \sum_i \alpha_i-i$. Figure~\ref{fig:one} shows ${\mathcal C}_1$ when $p=2$ and $m=3$. \begin{figure}[htb] $$\epsfxsize=1.8in \epsfbox{fig1.eps}$$ \caption{${\mathcal C}_1$\label{fig:one}.} \end{figure} Given $\alpha^{(a)}\in {\mathcal C}_q$, define the {\it quantum Schubert variety} $$ Z_{\alpha^{(a)}}\quad :=\quad \{z=(z_{\beta^{(b)}})\in {\mathcal K}_q\mid z_{\beta^{(b)}}=0\mbox{ if }\ \beta^{(b)}\not\leq \alpha^{(a)} \}\ . $$ Let ${\mathcal H}_{\alpha^{(a)}}$ be the hyperplane defined by $z_{\alpha^{(a)}}=0$. The main technical result we use is the following. \begin{prop}[\cite{RRW96,RRW98}]\label{prop:RRW} Let $\alpha^{(a)}\in{\mathcal C}_q$. Then \begin{enumerate} \item[(i)] $Z_{\alpha^{(a)}}$ is an irreducible subvariety of $\,{\mathcal K}_q$ of dimension $|\alpha^{(a)}|$. \item[(ii)] The intersection of $\,Z_{\alpha^{(a)}}$ and ${\mathcal H}_{\alpha^{(a)}}$ is generically transverse, and $$ Z_{\alpha^{(a)}}\cap {\mathcal H}_{\alpha^{(a)}}\ =\ \bigcup_{\beta^{(b)}\lessdot\alpha^{(a)}} Z_{\beta^{(b)}}\,. $$ \end{enumerate} \end{prop} Another proof of (ii) is given in~\cite{SS_SAGBI}, which shows (ii) is an ideal-theoretic equality. From (ii) and B\'ezout's theorem, we obtain the following recursive formula for the degree of $Z_{\alpha^{(a)}}$ $$ \deg Z_{\alpha^{(a)}}\ =\ \sum_{\beta^{(b)}\lessdot\alpha^{(a)}} \deg Z_{\beta^{(b)}}\,. $$ Since the minimal quantum Schubert variety is a point, we deduce the main result of~\cite{RRW98}: \begin{cor} The degree $\delta$ of $\,{\mathcal K}_q$ is the number of maximal chains in the poset ${\mathcal C}_q$. \end{cor} Closed formulas are given for $\delta$ in~\cite{RRW96,RRW98}, the source of the formula~(\ref{dqmp}), as well as the number $\deg Z_{\alpha^{(a)}}$ of maximal chains below $\alpha^{(a)}$. \section{Proof of Theorem~\ref{thm:real}} Let $L(s,t)$ be the $m$-plane osculating the parameterized rational normal curve $$ \gamma\ :\ (s,t)\in{\mathbb P}^1\ \longmapsto\ (s^{n-1},\,ts^{n-2},\,\ldots,\,t^{n-2}s,\,t^{n-1})\in{\mathbb P}^{n-1} $$ at the point $\gamma(s,t)$. Then $L(s,t)$ is the row space of the $m\times n$ matrix of forms with rows $\gamma(s,t),\gamma'(s,t),\ldots,\gamma^{(m-1)}(s,t)$, the derivative taken with respect to the parameter $t$. Write $L(s,t)$ for this matrix. For $\alpha\in\binom{[n]}{p}$, the maximal minor of $L(s,t)$ complementary to $\alpha$ is $(-1)^{|\alpha|}s^{\binom{m}{2}}l_\alpha s^{|\alpha|}t^{mp-|\alpha|}$, where $|\alpha|:=\sum_i \alpha_i-i$ and $(-1)^{|\alpha|}l_\alpha$ is the corresponding maximal minor of $L(1,1)$. Let ${\mathcal H}(s,t)$ be the pencil of hyperplanes given by the linear form $$ \Lambda(s,t)\ :=\ \sum_{\alpha^{(a)}\in{\mathcal C}_q} z_{\alpha^{(a)}} l_\alpha s^{|\alpha^{(a)}|}t^{N-|\alpha^{(a)}|}\,. $$ Let $M$ be a matrix representing a curve in ${\mathcal M}_q$. Then $$ \det\left[\begin{array}{c} L(s,t)\\M(s^n,t^n)\rule{0pt}{16pt}\end{array}\right]\ =\ s^{\binom{m}{2}} \sum_{\alpha^{(a)}\in{\mathcal C}_q} z_{\alpha^{(a)}} s^{an}t^{(q-a)n} l_\alpha s^{|\alpha|}t^{mp-|\alpha|}\ =\ s^{\binom{m}{2}}\Lambda(s,t)\,. $$ Thus ${\mathcal M}_q\cap {\mathcal H}(s,t)$ consists of all maps $M:{\mathbb P}^1\rightarrow {\bf G}$ of degree $q$ which satisfy $M(s^n,t^n)\cap L(s,t)\neq \{0\}$. Theorem~1 is a consequence of the following two theorems. \begin{thm}\label{thm:transverse} There exist positive real numbers $t_1,\ldots,t_N$ such that for any $\alpha^{(a)}\in{\mathcal C}_q$, the intersection $$ Z_{\alpha^{(a)}}\cap{\mathcal H}(1,t_1)\cap\cdots\cap {\mathcal H}(1,t_{|\alpha^{(a)}|}) $$ is transverse with all points of intersection real. \end{thm} \begin{thm}\label{thm:proper} If $t_1,\ldots,t_k\in {\mathbb C}$ are distinct, then for any $\alpha^{(a)}\in{\mathcal C}_q$, the intersection \begin{equation}\label{eq:proper} Z_{\alpha^{(a)}}\cap{\mathcal H}(1,t_1)\cap\cdots\cap{\mathcal H}(1,t_k) \end{equation} is proper in that it has dimension $|\alpha^{(a)}|-k$. \end{thm} \noindent{\bf Proof of Theorem~\ref{thm:real}. } By Theorem~\ref{thm:transverse}, there exist positive real numbers $t_1,\ldots,t_N$ (necessarily distinct) so that the intersection \begin{equation}\label{eq:int} {\mathcal K}_q \cap {\mathcal H}(1,t_1)\cap\cdots\cap{\mathcal H}(1,t_N) \end{equation} is transverse and consists of exactly $\delta$ real points. We show all these points lie in ${\mathcal M}_q$, and thus are maps $M:{\mathbb P}^1\rightarrow {\bf G}$ of degree $q$ satisfying $M(1,t_i^n)\cap L(1,t_i)\neq\{0\}$ for $i=1,\ldots,N$, which proves Theorem~\ref{thm:real}. Recall the map $\pi:{\mathbb P}^1\times{\mathcal K}_{q-1}\rightarrow {\mathcal K}_q$~(\ref{eq:pi}) whose image is the complement of ${\mathcal M}_q$ in ${\mathcal K}_q$. Then \begin{eqnarray*} \pi^*{\mathcal H}(s,t)&=& \sum_{\alpha^{(a)}\in{\mathcal C}_q} (Ax_{\alpha^{(a)}}-Bx_{\alpha^{(a-1)}}) l_\alpha s^{|\alpha^{(a)}|}t^{N-|\alpha^{(a)}|}\\ &=& (At^n-Bs^n)\sum_{\alpha^{(a)}\in{\mathcal C}_{q-1}} x_{\alpha^{(a)}}l_\alpha s^{|\alpha^{(a)}|}t^{N-n-|\alpha^{(a)}|}\,. \end{eqnarray*} Hence, if ${\mathcal H}'(s,t)$ is the pencil of hyperplanes in the Pl\"ucker space of ${\mathcal K}_{q-1}$ defining the locus of $M\in{\mathcal M}_{q-1}$ satisfying $M(s^n,t^n)\cap L(s,t)\neq \{0\}$, then $$ \pi^*{\mathcal H}(s,t)\ =\ (At^n-Bs^n){\mathcal H}'(s,t)\,. $$ Thus any point in~(\ref{eq:int}) not in ${\mathcal M}_q$ is the image of a point $([A,B],M)$ in ${\mathbb P}^1\times{\mathcal K}_{q-1}$ satisfying $\pi^*{\mathcal H}(1,t_i)=(At_i^n-B){\mathcal H}'(1,t_i)$ for each $i=1,\ldots,N$. As the $t_i$ are positive and distinct, such a point can only satisfy $At_i^n-B=0$ for one $i$. Thus $M\in{\mathcal K}_{q-1}$ lies in at least $N-1$ of the hyperplanes ${\mathcal H}'(1,t_i)$. Since $N-1$ exceeds the dimension $N-n$ of ${\mathcal K}_{q-1}$, there are no such points $M\in{\mathcal K}_{q-1}$, by Theorem~\ref{thm:proper} for maps of degree $q-1$. \qed\bigskip \noindent{\bf Proof of Theorem~\ref{thm:proper}. } For any $t_1,\ldots,t_k$, the intersection~(\ref{eq:proper}) has dimension at least $|\alpha^{(a)}|-k$. We show it has at most this dimension, if $t_1,\ldots,t_k$ are distinct. Suppose $k=|\alpha^{(a)}|+1$ and let $z\in Z_{\alpha^{(a)}}$. Then $z_{\beta^{(b)}}=0$ if $\beta^{(b)}\not\leq\alpha^{(a)}$ and so the form $\Lambda(1,t)(z)$ defining ${\mathcal H}(1,t)$ is divisible by $t^{N-|\alpha^{(a)}|}$ with quotient $$ \sum_{\beta^{(b)}\leq\alpha^{(a)}} z_{\beta^{(b)}}l_\beta\, t^{|\alpha^{(a)}|-|\beta^{(b)}|}\,. $$ This is a non-zero polynomial in $t$ of degree at most $|\alpha^{(a)}|$ and thus it vanishes for at most $|\alpha^{(a)}|$ distinct $t$. It follows that~(\ref{eq:proper}) is empty for $k>|\alpha^{(a)}|$. If $k\leq |\alpha^{(a)}|$ and $t_1,\ldots,t_k$ are distinct, but~(\ref{eq:proper}) has dimension exceeding $|\alpha^{(a)}|-k$, then completing $t_1,\ldots,t_k$ to a set of distinct numbers $t_1,\ldots,t_{|\alpha^{(a)}|+1}$ would give a non-empty intersection in~(\ref{eq:proper}), a contradiction. \qed\bigskip \noindent{\bf Proof of Theorem~\ref{thm:transverse}. } We construct the sequence $t_i$ inductively. If we let $\alpha=1<2<\cdots<p-1<p+1$, then $ Z_{\alpha^{(0)}}$ is a line. Indeed, it is isomorphic to the set of $p$-planes containing a fixed $(p-1)$-plane and lying in a fixed $(p+1)$-plane. By Theorem~\ref{thm:proper}, $Z_{\alpha^{(0)}}\cap{\mathcal H}(1,t)$ is then a single, necessarily real, point, for any real number $t$. Let $t_1$ be any positive real number. Suppose we have positive real numbers $t_1,\ldots,t_k$ with the property that for any $\beta^{(b)}$ with $|\beta^{(b)}|\leq k$, $$ Z_{\beta^{(b)}}\cap{\mathcal H}(1,t_1)\cap\cdots\cap {\mathcal H}(1,t_{|\beta^{(b)}|}) $$ is transverse with all points of intersection real. Let $\alpha^{(a)}$ be an index with $|\alpha^{(a)}|=k+1$ and consider the 1-parameter family ${\mathcal Z}(t)$ of schemes defined for $t\neq 0$ by $ Z_{\alpha^{(a)}}\cap{\mathcal H}(1,t)$. For $t\neq 0$, if we restrict the form $\Lambda(1,t)$ to $z\in Z_{\alpha^{(a)}}$, then, after dividing out $t^{N-|\alpha^{(a)}|}$, we obtain $$ z_{\alpha^{(a)}} + \sum_{\beta^{(b)}<\alpha^{(a)}} z_{\beta^{(b)}}l_\beta\, t^{|\alpha^{(a)}|-|\beta^{(b)}|}\,. $$ Thus ${\mathcal Z}(0)$ is $$ Z_{\alpha^{(a)}}\cap {\mathcal H}_{\alpha^{(a)}}\ =\ \bigcup_{\beta^{(b)}\lessdot\alpha^{(a)}} Z_{(\beta,d)}\,, $$ by Proposition~\ref{prop:RRW} (ii). \noindent{\bf Claim:} The cycle $$ {\mathcal Z}(0)\cap{\mathcal H}(1,t_1)\cap\cdots\cap{\mathcal H}(1,t_k) $$ is free of multiplicities. If not, then there are two components $ Z_{\beta^{(b)}}$ and $ Z_{\gamma^{(c)}}$ of ${\mathcal Z}(0)$ such that $$ Z_{\beta^{(b)}} \cap Z_{\gamma^{(c)}} \cap {\mathcal H}(1,t_1)\cap \cdots\cap {\mathcal H}(1,t_k) $$ is nonempty. But this contradicts Theorem~\ref{thm:proper}, as $Z_{\beta^{(b)}} \cap Z_{\gamma^{(c)}}=Z_{\delta^{(d)}}$, where $\delta^{(d)}$ is the greatest lower bound of $\beta^{(b)}$ and $\gamma^{(c)}$ in ${\mathcal C}_q$, and so $\dim Z_{\delta^{(d)}}<\dim Z_{\beta^{(b)}}=k$. From the claim, there is an $\epsilon_{\alpha^{(a)}}>0$ such that if $0\leq t\leq \epsilon_{\alpha^{(a)}}$, then $$ {\mathcal Z}(t)\cap{\mathcal H}(1,t_1)\cap\cdots\cap{\mathcal H}(1,t_k) $$ is transverse with all points of intersection real. Set $$ t_{k+1}\ :=\ \min\{\epsilon_{\alpha^{(a)}} : |\alpha^{(a)}|=k+1\}\,. \qquad\qed $$ \section{Further Remarks} From our proof of Theorem~\ref{thm:transverse}, we obtain a rather precise choice of $s_i$ and $L_i$ in the enumerative problem which give only real maps. By $\forall t_1\gg t_2\gg\cdots\gg t_N>0$, we mean $$ \forall t_1>0\ \ \exists \epsilon_2>0\ \ \forall \epsilon_2>t_2>0\ \ \cdots\ \ \exists \epsilon_N>0\ \ \forall \epsilon_N>t_N>0\, . $$ \begin{cor} $\forall t_1\gg t_2\gg\cdots\gg t_N>0$, each of the $\delta$ maps $M:{\mathbb P}^1\rightarrow{\bf G}$ of degree $q$ which satisfy $M(1,t_i)\cap L(1,t_i^{1/n})\neq \{0\}$ for $i=1,\ldots,N$ are real. \end{cor} When $q=0$, there is substantial evidence~\cite{Sottile_shapiro} that this choice of $t_1,\ldots,t_N$ is too restrictive. B.~Shapiro and M.~Shapiro have the following conjecture: \medskip \noindent{\bf Conjecture. } {\it Suppose $q=0$. Then for generic real numbers $\,t_1,\ldots,t_{mp}$ all of the finitely many $p$-planes $H$ which satisfy $H\cap L(1,t_i)\neq \{0\}$ are real. }\medskip In contrast, when $q>0$, the restriction $\forall t_1\gg t_2\gg\cdots\gg t_N>0$ is necessary. We observe this in the case when $q=1$, $p=m=2$, so $N=8$ and $\delta=8$. That is, for parameterized curves of degree 1 in the Grassmannian of 2-planes in ${\mathbb C}^4$. Here, the choice of $t_i=i$ in~(\ref{eq:int}) gives no real maps, while the choice $t_i=i^6$ gives 8 real maps. We briefly describe that calculation. There are 12 Pl\"ucker coordinates $z_{ij^{(a)}}$ for $1\leq i<j\leq 4$ and $a=0,1$. If we let $f_{ij}:=tz_{ij^{(0)}}+sz_{ij^{(1)}}$, then $$ f_{14}f_{23} - f_{13}f_{24} + f_{12}f_{34}\ =\ 0\,, $$ as $f_{ij}(s,t)\in{\bf G}$ for all $s,t$. The coefficients of $t^2$, $st$, and $s^2$ in this expression give three quadratic relations among the $z_{ij^{(a)}}$: $$ \begin{array}{c} z_{14^{(0)}}z_{23^{(0)}}\ -\ z_{13^{(0)}}z_{24^{(0)}}\ +\ z_{12^{(0)}}z_{34^{(0)}}\,,\\ z_{12^{(1)}}z_{34^{(0)}}\ -\ z_{13^{(1)}}z_{24^{(0)}}\ +\ z_{14^{(1)}}z_{23^{(0)}} \rule{0pt}{15pt} + z_{23^{(1)}}z_{14^{(0)}}\ -\ z_{24^{(1)}}z_{13^{(0)}}\ +\ z_{34^{(1)}}z_{12^{(0)}}\,,\\ z_{14^{(1)}}z_{23^{(1)}}\ -\ \rule{0pt}{15pt} z_{13^{(1)}}z_{24^{(1)}}\ +\ z_{12^{(1)}}z_{34^{(1)}}\,, \end{array} $$ and these constitute a Gr\"obner basis for the homogeneous ideal of ${\mathcal K}_1$\cite{SS_SAGBI}. Here, the form $\Lambda$ is $$ \begin{array}{c} {\displaystyle \ \ t^8z_{12^{(0)}}-2t^7z_{13^{(0)}}+t^6z_{14^{(0)}}+3t^6z_{23^{(0)}} -2t^5z_{24^{(0)}}+t^4z_{34^{(0)}}}\\ +\,t^4z_{12^{(1)}}-2t^3z_{13^{(1)}}+t^2z_{14^{(1)}}+3t^2z_{23^{(1)}} -2t\ \, z_{24^{(1)}}+\ \ z_{34^{(1)}}\,. \end{array} $$ We set $z_{34^{(1)}}=1$ and work in local coordinates. Then the ideal generated by the 3 quadratic equations and 8 linear relations $\Lambda(t_i)$ for $i=1,\ldots,8$ defines the 8 solutions to~(\ref{eq:int}). We used Maple V.5 to generate these equations and then compute a univariate polynomial in the ideal, which had degree 8. This polynomial had no real solutions when $t_i=i$, but all 8 were real when $t_i=i^6$. (Elimination theory guarantees that the number of real solutions equals the number of real roots of the eliminant.) \medskip We describe how the enumerative problem~(\ref{eq:enumerative}) arises in systems theory (see also~\cite{Byrnes}). A physical system (eg.~a mechanical linkage) with $m$ inputs and $p$ measured outputs whose evolution is governed by a system of linear differential equations is modeled by a $m\times n$-matrix $L(s)$ of real univariate polynomials. The largest degree of a maximal minor of this matrix is the MacMillan degree, $r$, of the evolution equation. Consider now controlling this linear system by output feedback with a dynamic compensator. That is, a $p$-input, $m$-output linear system $M$ is used to couple the $m$ inputs of the system $L$ to its $p$ outputs. The resulting closed system has characteristic polynomial $$ \varphi(s)\ :=\ \left[\begin{array}{c}L(s)\\M(s)\end{array}\right]\,, $$ and the roots of $\varphi$ are the natural frequencies or {\it poles} of the closed system. The dynamic pole assignment problem asks, given a system $L(s)$ and a desired characteristic polynomial $\varphi$, can one find a (real) compensator $M(s)$ of MacMillan degree $q$ so that the resulting closed system has characteristic polynomial $\varphi$? That is, if $s_1,\ldots,s_{r+p}$ are the roots of $\varphi$, which $M\in{\mathcal M}_q$ satisfy $$ \det\left[\begin{array}{c}L(s_i)\\M(s_i)\end{array}\right]\ =\ 0, \qquad\mbox{for }i=1,2,\ldots,r+p\,? $$ In the critical case when $r+q=mp+qn$, this is an instance of the enumerative problem~(\ref{eq:enumerative}). When the degree $\delta$ is odd, then for a real system $L$ and a real characteristic polynomial $\varphi$, there will be at least one real dynamic compensator. Part of the motivation for~\cite{RRW96} was to obtain a formula for $\delta$ from which its parity could be deduced for different values of $q,m$, and $p$. From this description, we see that the choice of planes $L_i$ that arise in the dynamic pole placement problem are $N=mp+qn$ points on a rational curve of degree $mp+(n-1)q$ in the Grassmannian of $m$-planes in ${\mathbb C}^n$. In contrast, the planes of Theorem~\ref{thm:transverse} (and hence of Theorem~\ref{thm:real}) arise as $N$ points on a rational curve of degree $mp$. Only when $q=0$ (the case of static compensators) is there any overlap. While our proof of Theorem~\ref{thm:real} owes much to systems theory, it has no direct implications for the problem of real dynamic output compensation. \medskip Our method of proof of Theorem~\ref{thm:real} (like that in~\cite{So99}) was inspired by the numerical Pieri homotopy algorithm of~\cite{HSS} for computing the solutions to~(\ref{eq:enumerative}) when $q=0$. Likewise, the explicit degenerations of intersections of the ${\mathcal H}(s,t)$ that we used, and more generally Proposition~\ref{prop:RRW} (ii), can be used to construct an optimal numerical homotopy algorithm for finding the solutions to~(\ref{eq:enumerative}). This is in exactly the same manner as the explicit degenerations of intersections of special Schubert varieties of~\cite{So97d} were used to construct the Pieri homotopy algorithm of~\cite{HSS}. \medskip We close with one open problem concerning the enumeration of rational curves on a Grassmannian. For a point $s\in{\mathbb P}^1$ and any Schubert variety $\Omega$ of ${\bf G}$, consider the quantum Schubert variety $\Omega(s)$ of curves $M\in{\mathcal M}_q$ satisfying $M(s)\in\Omega$. The quantum Schubert calculus gives algorithms to compute the number of curves $M\in{\mathcal M}_q$ which lie in the intersection of an appropriate number of these $\Omega(s)$, and we ask when it is possible to have all solutions real. A modification of the proof of Theorem~\ref{thm:transverse} shows that this is the case when all except possibly 2 are hypersurface Schubert varieties. In every case we have been able to compute, all solutions may be real.
\section{Introduction} \label{secintro} \indent At a normal metal - superconductor (NS) junction, electrons incident from the normal metal can be scattered into time reversed electrons (holes) by the pairing potential. This conversion process is known as Andreev reflection.~\cite{rAndreev} When the NS junction carries an electrical current, Andreev reflection is accompanied by the conversion of normal current to supercurrent.~\cite{rKum,rMathews} This supercurrent flow modifies the current-voltage relation in NS and NSN junctions~\cite{rSanSol}~\cite{rSanSol2}~\cite{rLamb}, superconducting wires~\cite{rLiFu,rBag1}, SNS junctions~\cite{rRie1}, and NS junctions with a supercurrent parallel to the NS interface~\cite{rHofKum}. In this paper we consider the current-voltage relation for NS junctions having a supercurrent flow perpendicular to the NS junction. \begin{figure}[htb] \centps{riedelfig1.eps}{45} \caption{Different types of NS Junctions. (a) NS point contact, (b) N-Narrow S- S, (c) N - Wider S- S. The wire width determines the number of conducting modes in the narrow segments ($M_N$ and $M_S$). The superfluid flow velocity $v_s$ cannot be neglected when $M_N \simeq M_S$.} \label{fig1DGeom} \end{figure} The superfluid flow present for the point contact NS junction in Fig.~\ref{fig1DGeom}(a) will have little effect on its I-V characteristic, since the current density inside the wide superconductor approaches zero. However, for the NS junctions shown in Fig.~\ref{fig1DGeom}(b)-(c) , the number of conducting modes in the superconducting wire ($M_S$) is comparable to to the number of conducting modes in the normal wire ($M_N$). Since the current density is not zero inside the superconductor, one cannot neglect the effect of a superfluid flow on the I-V characteristics of the NS junctions shown in Fig.~\ref{fig1DGeom}(b)-(c). Since the superfluid flow strongly modifies the dispersion relationship of the superconductor when the ratio of the number of conducting modes $\alpha = M_S / M_N$ is of order one, including such a superfluid flow will influence the I-V characteristic of NS junctions. Since the number of conducting modes is roughly proportional to the width of the conductor, namely $\alpha \simeq W_S / W_N$, we can vary the ratio of conducting modes by varying the width of the superconductor $W_S$. Blonder, Tinkham and Klapwijk (BTK) showed that the point contact NS junction in Fig.~\ref{fig1DGeom}(a) carries a larger current than a normal metal point contact junction~\cite{rBTK}. This 'excess' current of is due to the presence of Andreev reflection, and has the value $I_{\rm exc} = (4/3)(2e\Delta/h)$ in a ballistic point-contact NS junction. By varying the width of the superconducting wire forming the NS junction in Fig.~\ref{fig1DGeom}(c), we determine in this paper how the excess current varies as a function of the transverse mode ratio $\alpha$. Using both an independent band model, and a second model which includes scattering between different lateral modes at the NS interface, we show the excess current in NS junction can be much larger than the BTK result. This enhancement of the excess current over the BTK value has also been noted in Ref.~\cite{rSanSol2}. More Andreev reflections can occur at higher voltages when a supercurrent flows perpendicular to the NS interface, accounting for an excess current larger than the BTK result. If the superconductor is narrower than about $W_S < (7/3) W_N$, the narrow region of the superconductor exceeds its critical current before allowing the maximum number of Andreev reflections. If the superconductor is much wider than $W_S >> (7/3) W_N$, there is too much geometrical dilution of the supercurrent for a significant superfluid flow to develop inside the narrowest region of the superconductor. We find a maximum excess current when the width of the superconductor is approximately $W_S = (7/3) W_N$. \section{Why Superfluid Flow Increases the Excess Current} \label{secwhymore} \indent In this section we give the simplest physical model which illustrates why the excess current in NS junctions can be larger than the point contact limit. To make our physical points, we construct a crude two fluid model, which does not obey electrical current conservation at every point in space. A fully self-consistent solution of the Bogoliubov-deGennes equations automatically ensures current conservation, eliminating the need for this ad-hoc two-fluid model, but requires more computational effort. The two-fluid model we develop only guarentees current conservation at the terminal contacts. A fully self-consistent solution of the BdG equations in a single mode junction ($\alpha = 1$), done in section \ref{secself1d}, gives the same results for the excess current as the two-fluid model. Viewing the electrical conduction in terms of this two fluid model, therefore, allows us to obtain a value of the superfluid velocity $v_s$ inside the superconducting contacts for each value of the bias voltage across the NS or NIS junction. We then use this value of the flow velocity $v_s$ to compute Andreev and normal reflection probabilities for each value of the voltage, and thus obtain the electrical current in a globally self-consistent manner. \subsection{Two Fluid Model} \label{twoflu} \indent Figure~\ref{figBand} shows the energy band diagram of an NS junction when the superconductor carries a finite supercurrent~\cite{rBag1}. In a transmission formalism, one must compute the electrical current operator for all incident quasi-particle states, and add them to obtain the total current. Ref.~\cite{rDatta1} evaluates the electrical current operator on the normal side of the NS junction in terms of particle current transmission and reflection probabilities. The derivation in section 3 of Ref.~\cite{rDatta1} is valid for a multiple moded NS junction subject to a superfluid flow. Ref.~\cite{rDatta1} recovers the well-known BTK current formula~\cite{rBTK}, namely (for zero temperature) \begin{equation} I = (2e/h) M_N \int^{eV}_{0} [1 - T_{Ne,Ne}(E) + T_{Nh,Ne}(E)] dE. \label{I_NT=0} \end{equation} In Eq.~(\ref{I_NT=0}) we use the notation of Ref.~\cite{rDatta1}, where the $T_{N \delta, N \beta}$ particle current reflection probabilities from the incident channel $(N \beta)$ to reflected channel $(N \delta)$. The indices $\beta, \delta = e \; {\rm or } \; h$ for electron-like or hole-like quasi-particles. We must compute the both the normal $T_{Ne,Ne}=R_N$ and Andreev $T_{Nh,Ne}=R_A$ reflection probabilities when the superconducting contact is subject to a superfluid flow. For simplicity we have taken the $M_N$ modes in the normal conductor to be both independent and to carry identical electrical currents. \begin{figure}[htb] \centps{riedelfig2.eps}{60} \caption{ Energy band diagram of an NS junction subject to a superfluid flow $v_S \ge 0$. The shifted energy bands in the superconductor cause Andreev reflection to occur at higher energies than without superfluid flow. The contacts inject electron-like (solid dots) and hole-like (open circles) quasi-particles as shown.} \label{figBand} \end{figure} In order for the superconductor to carry a finite supercurrent, the form of the order parameter $\Delta(x)$ inside the superconductor must be generalized to $\Delta(x) = |\Delta| e^{2iqx}$. The superfluid velocity is $v_{s}= \hbar q / m$. If we consider solutions accurate within the Andreev approximation ($|E| \ll \mu$), we can approximate the dispersion relation near the Fermi level as being rigidly shifted in energy as~\cite{rBag1} \begin{eqnarray} \frac{\hbar^{2}k^{2}}{2m} - \mu \simeq \pm \sqrt{(E \mp |\Delta| (q/q_d))^{2} - |\Delta|^{2}} \; . \label{shiftE} \end{eqnarray} Here $v_d = \hbar q_d / m = |\Delta| / \hbar k_F$ is the Landau depairing velocity.~\cite{rLan} The electrical current is given by $I_C \simeq e n v_s M_S$ with $n = 2(k_F/\pi)$ the electron density per mode and $M_S$ the number of (equivalent, for simplicity) conducting modes in the superconductor~\cite{rLiFu}. We can rewrite the supercurrent $I_C$ in terms the depairing velocity as \begin{equation} I_C \simeq (4e|\Delta|/h) (v_s/v_d) M_S . \label{I_C} \end{equation} At zero temperature, the critical current phase boundary occurs when $v_s = v_d$. We must therefore maintain $v_s < v_d$ to preserve the superconducting order parameter $|\Delta| \ne 0$. Inside the superconductor, the electrical current is often argued to be composed of a `quasi-particle' and `condensate' contribution~\cite{rBTK,rPethik}. To break Eq.~(\ref{I_NT=0}) down into these two contributions to the current we use the sum rule~\cite{rBTK} \begin{equation} 1 = T_{Ne,Ne}(E) + T_{Nh,Ne}(E) + T_{Se,Ne}(E) + T_{Sh,Ne}(E). \label{Pconserv} \end{equation} Equation~(\ref{Pconserv}) states that the normal ($T_{Se,Ne}=T_{N}$) and Andreev ($T_{Sh,Ne}=T_{A}$) particle current transmission coefficients into the superconductor conserve the total number of quasi-particles. Combining Eqs.~(\ref{I_NT=0}) and (\ref{Pconserv}), the electrical current inside the superconductor at zero temperature is \begin{equation} I = I_{QP} + I_A. \label{I_S} \end{equation} We identify the portion of the electrical current due to `quasi-particle' injection as \begin{equation} I_{QP} = (2e/h) M_N \int^{eV}_{0} [T_{Se,Ne}(E) - T_{Sh,Ne}(E)] dE, \label{I_QP} \end{equation} and the `Andreev' portion of the current $I_A$ as \begin{equation} I_A = (2e/h) M_N \int^{eV}_{0} 2 [T_{Nh,Ne}(E)+ T_{Sh,Ne}(E) ] dE . \label{2And} \end{equation} The `Andreev' current in Eq.~(\ref{2And}) equals twice the sum of all Andreev processes. Equations~(\ref{I_S})-(\ref{2And}) give the same current as the BTK expression from Eq.~(\ref{I_NT=0}), since both are simply the current operator evaluated inside the normal metal. The key physical element in our `two-fluid' model is that we require that both Eqs.~(\ref{I_C}) and (\ref{2And}) for the `condensate' current $I_C$ and the `Andreev' current $I_A$ must be equal. We examine this assumption more rigorously in the Appendix, by evaluating the electrical current operator inside the superconductor. In this section we make a plausibility argument for equating $I_A$ from Eq.~(\ref{2And}) with the `condensate current' $I_C$ from Eq.~(\ref{I_C}), or the `current of Cooper pairs'. Eq.~(\ref{2And}) expresses the condensate current in terms of probabilities for Andreev reflection and Andreev transmisison of an electron incident from the normal metal. Conversely, Eq.~(\ref{I_C}) expresses the condensate current in terms of the superfluid velocity. That an incident electron and a reflected hole on the normal metal side requires a Cooper pair to move off into the superconductor (to preserve electrical charge conservation) is well known~\cite{rBTK}. Similarly, Andreev transmission of an electron also requires a Cooper pair flow inside the superconductor. Andreev reflection and Andreev transmission of electrons incident from the normal metal require Cooper pairs flow away from the NS interface (in the same direction) for both processes. Equation~(\ref{2And}) embodies this physical reasoning. Similarly, identifying Eq.~(\ref{I_QP}) as the `quasi-particle' contribution to the electrical current is also quite natural, being proportional to the electrical current operator for an electron injected from the normal metal (evaluated on the superconducting side). The `two fluid' picture here requires Eqs.~(\ref{I_C}) and (\ref{2And}) be equal be satisfied in order to guarentee global current conservation. Once the junction geometry and applied voltage is specified, the only free parameter in Eqs.~(\ref{I_C}) and (\ref{2And}) is the superfluid velocity $v_s=\hbar q/m$. Since the quasi-particle transmission and reflection coefficients themselves depend on $v_s$, equating Eqs.~(\ref{I_C}) and (\ref{2And}) is then a globally self-consistent procedure for determining the superfluid flow velocity $v_s$. Using this value for $v_s$, one then uses either Eq.~(\ref{I_NT=0}) or (\ref{I_S}) to find the terminal currents for each value of the bias voltage $V$. \subsection{Two-Fluid Approximation for Ballistic NS Junction} \label{balns} \indent We now restrict our attention to ballistic NS junctions, and approximate the energy bands near the Fermi level as simply rigidly shifted in energy by an amount $\pm |\Delta| (v_s/v_d)$, as determined from Eq.~(\ref{shiftE}). When the superconductor carries a finite supercurrent, we can then obtain the Andreev reflection coefficient from \begin{equation} T_{Nh,Ne}(E) = R_A^0(E - |\Delta| (v_s/v_d)). \label{RAshift} \end{equation} Here $R_A^0(E)$ is the Andreev reflection coefficient found by BTK~\cite{rBTK} when the superfluid flow is zero ($v_s=0$), namely \begin{equation} R_A^0(E)= \left\{ \matrix{ 1 & |E| \leq |\Delta| \cr \left( \frac{\displaystyle E - \sqrt{E^{2} - |\Delta|^{2}}} {\displaystyle E + \sqrt{E^{2} - |\Delta|^{2}}} \right) & |E| \geq |\Delta| } \right. . \label{RA0} \end{equation} This rigid shift in the Andreev reflection coefficient, corresponding to the rigid shift in the energy bands near the Fermi level, is shown in Fig.~\ref{figAnd}. Simply shifting the reflection coefficients in energy is not a valid approximation when a tunnel barrier is present at the NS interface~\cite{rRieThes}, as also noted in Fig.~\ref{figAnd}. Since the differential conductance of an NS junction is \begin{equation} \left. \frac{dI}{dV} \right|_V = \frac{2e}{h} [ 1 + R_A (E = eV) - R_N(E = eV) ] , \end{equation} a differential conductance measurement producing a larger than expected energy gap could point to significant superfluid flow in the junction. Shifting the Andreev reflection coefficient in energy makes it possible for the `negative energy' Andreev reflections, i.e. the Andreev reflection probabilities having $E<0$ in Eqs.~(\ref{RA0}) to contribute to the excess electrical current. These `negative energy' Andreev reflections are shown as the additional area under the Andreev reflection probability for $E>0$ when $v_s > 0$ in Fig.~\ref{figAnd}, At zero temperature, Eqs.~(\ref{I_NT=0}) and (\ref{RAshift}) give \begin{equation} I = (2e/h) M_N \int^{eV}_{0} [1 + R_{A}^0(E - |\Delta| (v_s/v_d))] dE. \label{INbalT=0} \end{equation} To determine the superfluid velocity $v_s$ in Eq.~(\ref{INbalT=0}), Eqs.~(\ref{I_C}) and (\ref{2And}) require \begin{equation} |\Delta| \alpha (v_s/v_d) = \int^{eV}_{0} R_{A}^0(E - |\Delta| (v_s/v_d)) dE. \label{selfconvs} \end{equation} Equation~(\ref{selfconvs}) is a self-consistent equation for the superfluid velocity $v_s$, and depends on the ratio of the number of conducting modes in the superconductor to the normal conductor $\alpha = M_S/M_N$. The largest possible excess current would occur if we could fix $v_s \to \infty$ in Eqs.~(\ref{INbalT=0}), and would give twice the BTK result of $I_{\rm exc}^{max} = 2 I_{\rm exc}^{BTK} = (8/3)(2e\Delta/h)$. However, this theoretical maximum excess current is not possible due to the constraint that the superfluid velocity must remain smaller than the depairing velocity. One must therefore discard any solution of Equation~(\ref{selfconvs}) giving $v_s > v_d$. \begin{figure}[htb] \centps{fig3.eps}{60} \caption{Andreev reflection probability for an electron incident on an NS interface when the superfluid velocity is zero (solid line) and when $v_S = v_d/2$ (dashed line). Superfluid flow simply shifts the Andreev reflection probability by an amount $\Delta (v_S/v_{d})$. If a tunnel barrier is placed at the NS interface (dotted line), there is no simple relation between the Andreev reflection probabilities with and without superfluid flow.} \label{figAnd} \end{figure} \subsection{Numerical Evaluation of Two-Fluid Formula} \label{numeval} \indent We plot the numerical solution of Eqs.~(\ref{INbalT=0})-(\ref{selfconvs}) for the total current through the NS junction as a function of the voltage in Fig.~\ref{figIV}. When the transverse mode ratio is small, namely when $\alpha \leq 7/3$ shown in Fig.~\ref{figIV}(a), there is a voltage above which current conservation requires that $v_s$ exceed the Landau depairing velocity of the superconductor. When $v_s \geq v_d$, the narrower superconducting region (between the large normal contact and the large superconducting reservoir) becomes a normal conductor. This collapse of the order parameter in the narrower superconducting wire continues until a new NS interface is formed where the narrow conductor meets the wide superconducting reservoir. Geometrical dilution of the supercurrent where the superconductor widens into a thermodynamic reservoir moves the NS interface so that a stable point contact junction is formed. Forcing the narrower superconducting region into the normal state therefore creates a point contact NS junction having $\alpha = \infty$, i.e. the BTK limit of an NS junction. Forcing the narrower superconducting region to become normal therefore forces the excess current to fall abruptly to the BTK limit shown in Fig.~\ref{figIV}(a). \begin{figure}[htb] \twofig{fig4a.eps}{fig4b.eps}{6.0} \caption{ (a) When the transverse mode ratio $\alpha < (7/3)$, the order parameter in the narrow superconducting wire can collapse, giving rise to discontinuities in the I-V relation. (b) When $\alpha > (7/3)$ no such discontinuities arise. The excess current is larger than the BTK result for $\alpha = 7/3$, shown in both (a) and (b). The I-V evolves smoothly into the BTK result for a point contact NS junction when $\alpha \to \infty$.} \label{figIV} \end{figure} An excess current larger than the BTK limit is also shown in Fig.~\ref{figIV}(a), due to the additional `negative energy' Andreev reflections. When the transverse mode ratio is larger, namely when $\alpha \geq 7/3$ shown in Fig.~\ref{figIV}(b), the superfluid velocity in the narrower superconductor is always less than the Landau depairing velocity. Consequently, no abrupt drops in the current occur for any value of the voltage. The excess current simply decreases gradually from its maximum value at $\alpha = 7/3$ to the BTK value for a NS point contact at $\alpha = \infty$. We plot the maximum excess current in a ballistic NS junction versus the mode ratio $\alpha$ in Figure~\ref{figExCur}. The excess current at any given voltage is defined as the difference between the current carried by the NS superconducting junction and a normal NN junction, namely $I_{ex}(V) = I_{NS}(V) - I_{NN}(V)$. The maximum excess current occurs at voltage for which $I_{ex} = {\rm Max} \; [ I_{ex}(V) ]$. For $\alpha < 4/3$ the excess current is given by the point contact value, namely the BTK result of $(4/3)(2e\Delta/h)$, and occurs at a voltage $V = \infty$. When $4/3 \leq \alpha \leq 7/3$ the excess current occurs at a finite voltage $V \leq \infty$, and is larger than the BTK result. As the transverse mode ratio increased above $\alpha \geq 7/3$, geometrical dilution of the supercurrent reduce the band tilting in the superconductor, reducing the maximum excess current of the junction. When the mode ratio is very large, so that $\alpha \to \infty$ we recover the point contact result for the excess current. \begin{figure}[htb] \centps{fig5.eps}{60} \caption{The excess current $I_{\rm exc}$ of a ballistic normal-superconducting wire at zero degrees Kelvin is maximum when the transverse mode ratio is $\alpha = (M_S/M_N) = 7/3$. $I_{\rm exc}$ approaches the point contact (BTK) value of $4/3 (2e\Delta/h)$ when $\alpha \to \infty$. Collapse of the order parameter in the narrow superconducting wire limits $I_{\rm exc}$ when $\alpha < 7/3$.} \label{figExCur} \end{figure} The circled dots in Figure~\ref{figExCur} show the results of a more realistic numerical calculation~\cite{rRieThes} for the excess current. As detailed in Ref.~\cite{rRieThes}, we permit interband scattering at the NS junction and allow different conducting modes in the superconductor to carry different amounts of supercurrent~\cite{rLiFu}. The general behavior for the excess current versus mode ratio $\alpha$ of this more realistic NS junction model is quite similar to our simplified model, except the maximum in the excess current is slightly smaller and shifted to a slightly larger transverse mode ratio. The slightly lower excess current arises because the higher lying modes have a smaller Fermi velocity, and therefore the Andreev reflection coefficients for the higher lying modes do not shift in energy as much as the lower lying modes. We conclude that this more realistic model, even though still a two-fluid type model which only globally conserves electrical current, confirms the essential features of our simpler independent mode calculation in this section. \subsection{Limiting Cases of Two-Fluid Formula} \label{limitcases} \indent We can understand how the excess current $I_{exc}$ depends on the mode ratio $\alpha = M_S / M_N$ in Figure~\ref{figExCur} by examining Eqs.~(\ref{INbalT=0})-(\ref{selfconvs}) in different limits. The integral of the shifted Andreev reflection probability in Eqs.~(\ref{INbalT=0})-(\ref{selfconvs}) can be done analytically to yield \begin{equation} \int^{eV}_{0} R_{A}^0(E - |\Delta| (v_s/v_d)) dE = eV \label{intRA01} \end{equation} when $eV \leq |\Delta| (1 + (v_s/v_d))$. In this limit, Eqs.~(\ref{selfconvs}) and (\ref{intRA01}) show the superfluid velocity increases linearly with bias voltage $V$. For larger biases, namely when $eV \geq |\Delta| (1 + (v_s/v_d))$, the superfluid velocity increases more slowly with voltage, as determined from \begin{eqnarray} & & \int^{eV}_{0} R_{A}^0(E - |\Delta| (v_s/v_d)) dE \nonumber \\ & = & |\Delta| (1 + (v_s/v_d)) + |\Delta| [ \frac{1}{3} - \frac{1}{2} e^{-\gamma} + \frac{1}{6} e^{-3 \gamma} ] . \label{intRA02} \end{eqnarray} The factor $\gamma$ in Eq.~(\ref{intRA02}) is \begin{equation} \gamma = \cosh^{-1}(\frac{eV}{|\Delta|} - \frac{v_s}{v_d}) . \label{gamma} \end{equation} The excess current we obtain from \begin{equation} I_{exc} = (2e/h) M_N \int^{eV}_{0} R_{A}^0(E - |\Delta| (v_s/v_d)) dE. \label{IexcT=0} \end{equation} Consider first the case where the narrower superconducting wire is not driven normal, so that $v_s < v_d$ for all values of voltage. In that case, the maximum excess current occurs when $V = \infty$, so that $\gamma = \infty$ in Eq.~(\ref{gamma}). Eqs.~(\ref{selfconvs}) and (\ref{intRA02}) for the superfluid velocity then reduce to $(\alpha - 1)(v_s/v_d) = 4/3$. The maximum allowed superfluid velocity, $(v_s/v_d) = 1$, then occurs for a transverse mode ratio of $\alpha = 7/3$. The excess current from Eqs.~(\ref{IexcT=0}) and (\ref{intRA02}) then becomes \begin{equation} I_{exc} = \left( \frac{2e |\Delta|}{h} M_N \right) \left( \frac{4}{3} \right) \left( \frac{\alpha}{\alpha-1} \right) , \label{IexcT=0high} \end{equation} when $\alpha \geq 7/3$. The excess current reaches its maximum value of $I_{exc} = (7/3) (2e |\Delta|/h) M_N$ when $\alpha \geq 7/3$ as shown in Figure~\ref{figExCur}. Taking the limit $\alpha \to \infty$ in Eq.~(\ref{IexcT=0high}) recovers the BTK result $I_{exc} = (4/3) (2e |\Delta|/h) M_N$. We can also obtain an analytical solution for the excess current using Eqs.~(\ref{selfconvs}) and (\ref{intRA01}) to determine the superfluid velocity as $(v_s/v_d) = eV / |\Delta| \alpha$. The excess current then follows from Eqs.~(\ref{IexcT=0}) and (\ref{intRA01}) as $I_{exc} = (2e |\Delta|/h) M_N |\Delta| \alpha (v_s/v_d)$. Using the maximum alllowed depairing velocity of $(v_s/v_d) = 1$ then gives \begin{equation} I_{exc} = \left( \frac{2e |\Delta|}{h} M_N \right) \alpha. \label{IexcT=0low} \end{equation} The range of allowed mode ratios $\alpha$ for which Eq.~(\ref{IexcT=0low}) is valid lie along the curve $(v_s/v_d) = 1 = eV / |\Delta| \alpha$. Furthermore, Eq.~(\ref{intRA01}) is only valid for $eV / |\Delta| \leq (1 + (v_s/v_d))$. Combining all these requirements restricts the mode ratio between $1 \leq \alpha \leq 2$. However, for $1 \leq \alpha \leq 4/3$ the order parameter in the narrower superconducting wire collapes before the excess current reaches the BTK value. Eq.~(\ref{IexcT=0low}) therefore describes the excess current between $4/3 \leq \alpha \leq 2$ as shown in Figure~\ref{figExCur}. \section{Exact Solution of a Single Mode NS Junction} \label{secself1d} \indent In this section we wish to evaluate several assumptions made in the two fluid model of section \ref{secwhymore}. To do this, we solve the BdG equation, together with the self-consistency requirement for the order parameter, in a single band (1D) model of the NS junction where $M_N = M_S = 1$ ($\alpha = 1$). This self-consistent solution allows us to demonstrate how the supercurrent flow develops naturally from a self-consistent solution of the NS junction under a voltage bias. We use this solution to verify the approximate (two-fluid) procedure we use to guarentee global current conservation in section \ref{secwhymore}. One might expect that conserving the current only globally certainly gives qualitatively correct answers for the I-V characteristic. However, since the Andreev reflection probability does not vary much (as a function of energy) if we allow the order parameter to reach its final self-consistent form, the two-fluid model also gives accurate quantitative estimates for the I-V relation. Our self-consistent solution of the BdG equations in this section verifies the main assumptions used in our two-fluid model (when $\alpha = 1$). When a voltage is applied to the ballistic NS junction junction, the magnitude of $\Delta(x)$ remains approximately constant inside the superconductor (at zero temperature). However, the order parameter phase varies approximately linearly inside the superconducting metal. There are essentially no additional quasi-particles injected into the single moded NS junction at zero temperature, so the total current is simply $I = e n v_s$. The slope of the phase is related to the supercurrent velocity as $d \phi/dx = 2q = 2 v_s m / \hbar $. The superfluid velocity is linearly related to the voltage as $I = (4e^2/h) V = e n v_s$. As voltage bias increases, the slope of the phase $d \phi /dx$ increases until the superfluid velocity reaches the Landau depairing velocity, $v_s = v_d$ (or $d \phi /dx = 1/\xi_0$). At this voltage $V= V_{c}$ the ordering parameter inside the narros superconductor collapses, and a stable point contact NS junction forms inside the wide superconducting reservoir. \subsection{Self-Consistent Solution Procedure for BdG Equation} \label{selfconproc} The motion of quasi-particles in our one band NS junction, including the superfluid flow inside in the superconductor, is determined from the 1D time independent Bogliobov-de Gennes~\cite{rBdG} (BdG) equation \begin{equation} \left( \matrix{ H(x)-\mu & \Delta(x) \cr \Delta^*(x) & -(H^*(x) - \mu ) \cr} \right) \left( \matrix{ u(x) \cr v(x) } \right) = E \left( \matrix{ u(x) \cr v(x)} \right) \label{eBdG} \end{equation} The one-electron Hamiltonian $H(x)$ in Eq.~(\ref{eBdG}) is \begin{equation} H(x) = -\frac{\hbar^2}{2m} \frac{d^{2}}{dx^{2}} + V(x) . \end{equation} The order parameter $\Delta(x)$ in Eq.~(\ref{eBdG}) is given by \begin{eqnarray} \Delta(x) & = & g(x) F(x) \nonumber \\ & = & g(x) \sum_{pn} v^{*}_{pn}(x) u_{pn}(x) f_{pn} \theta(|E_{pn}| - \hbar \omega_D), \label{eCond} \end{eqnarray} where $g(x)$ is the electron-phonon interaction strength at each point and $F(x)$ is the pair correlation function. (Although we assume $g(x)$ is local, in reality it is spread over a correlation distance $v_F/\omega_D$.) The index $p$ in Eq.~(\ref{eCond}) denotes the lead from which the scattering state originates, namely $p=N$ is the left lead and $p=S$ the right lead. The index $n$ in Eq.~(\ref{eCond}) denotes the good quantum numbers in the lead, namely $n = (k,\beta)$ where $k$ is the wavenumber and $\beta = (e,h)$ the electron-like and hole-like states. The sum in Eq.~(\ref{eCond}) runs over states injected from the leads, including both positive ($E_n > 0$) and negative energies ($E_n < 0$). The coherence factors $u(x)$ and $v(x)$ in Eq.~(\ref{eCond}) are functions of the order parameter $\Delta(x)$ through Eq.~(\ref{eBdG}). In this section we show the self-consistent solutions of Eq.~(\ref{eBdG}) and (\ref{eCond}) for a voltage-biased NS junction. Details of the self-consistent solution procedure are given in Ref.~\cite{rRie1}. To solve the order parameter self-consistently, we first assure an initial or zeroth order guess $\Delta_0(x)$ for the order parameter. We then divide the one dimensional space is into differential elements, where the magnitude of the order parameter superfluid velocity are constant in each section. We match the wavefunctions and their derivatives at each interface to obtain the zeroth order wavefunctions $u_0(x)$ and $v_0(x)$. The first iteration for the order parameter $\Delta_1(x)$ we then obtain from Eq.~(\ref{eCond}) using the zeroth order wavefunctions $u_0(x)$ and $v_0(x)$, etc. The zeroth order guess for the order parameter $\Delta_0(x)$ can either be constant, i.e. $\Delta_0(x) = \Delta$, or it can contain a superfluid flow, i.e. $\Delta_0(x) = \Delta e^{2iqx}$. Given the same electrical current flow, either initial guess for the order parameter converges to the same final answer. The voltage bias across the NS junction is a boundary condition which determines the Fermi occupation probabilities $f_{pn}$ in Eq.~(\ref{eCond}). The occupation probability $f_{pn}$ of a scattering state $(p,n)$ which originates inside the normal or superconducting reservoir $p$, is the same for holes and electrons when the applied bias is zero ($V=0$). Under a voltage bias, however, electrons in the normal metal are occupied up to an energy $\mu+eV$, while holes are occupied up to an energy $\mu-eV$.~\cite{rDatta1} These different Fermi factors electron-like and hole-like quasi-particles injected from the normal metal are shown schematically in Fig.~\ref{figBand}. The unequal occupation probabilities for holes and electrons injected from the normal metal causes these two classes of scattering states to contribute differently to the sum in Eq.~(\ref{eCond}) under an applied bias. We can write the Fermi factors as \begin{equation} f_{pn} = f_{p \beta} = f(E - eV_{p \beta} ) \; , \label{fermi} \end{equation} where $f(E) = 1/[1+{\rm exp} \left( E / k_B T \right)]$. Here $eV_{p \beta}$ is effective biasing voltage (or effective electrochemical potential) applied to the $(p \beta)$th lead, namely~\cite{rDatta1} \begin{equation} V_{Ne} = V \; , \label{vefactors} \end{equation} \begin{equation} V_{Nh} = - V \; . \label{vhfactors} \end{equation} In this paper the superconducting leads are grounded so that $V_{S \beta} = 0$. To obtain the electrical current $I(x)$, we do not invoke any ad hoc `source term' as done in Ref.~\cite{rBTK}, but instead simply evaluate the electrical current operator for a scattering state originating in lead $q$ (having quantum number $n$) and terminating in lead $p$, namely \begin{equation} I_{q} = \sum_{pn} \left( J_{u} + J_{v} \right)_{q ; pn} f_{pn} - \sum_{p n} \left( J_{v} \right)_{q ; p n} \; . \label{eCur} \end{equation} The $J_{u}$ and $J_{v}$ are Schr\"{o}dinger currents associated with the waves $u$ and $v$, namely $J_{u}=(e\hbar/m) {\rm Im}\{u^{*}(x) \nabla u(x)\}$ and $J_{v}=(e\hbar/m) {\rm Im}\{v^{*}(x) \nabla v(x)\}$. The `vacuum current' due to the filled hole band is argued in Ref.~\cite{rDatta1} to be zero, namely $\sum_{q n} (J_{v})_{p ; q n} = 0$, as we have confirmed for the NS junction. Solving Eq.~(\ref{eBdG}) together with Eq.~(\ref{eCond}) guarentees electrical current conservation~\cite{rKum,rMathews},\cite{rBag1},\cite{rSols,rFurusaki3}, even when the superconductor is far from equilibrium. A proof of this statement for NS junctions follows from generalizing the discussion in Appendix B of Ref.~\cite{rBag1} to the nonequilibrium case. If the nonequilibrium system involves two superconductors at different biases~\cite{rHurd}, current conservation is more complex and $\sum_{q n} (J_{v})_{p ; q n} \ne 0$. \subsection{Order Parameter Phase} \label{opphase} Consider first the NS junction, where the coupling constant $g(x)$ is \begin{equation} g(x) = \left\{ \matrix{ g_R & x > 0 \cr 0 & x < 0 } \right. \end{equation} We choose $g_R$ and $\omega_D$ so that the critical temperature of the right superconductor is $T_c = 6.6$K. Our initial guess for the order parameter we take to be \begin{equation} \Delta_0(x) = \left\{ \matrix{ \Delta & x > 0 \cr 0 & x < 0 } \right. \label{delta0} \end{equation} We therefore do not force a superfluid flow inside the superconductor from our zeroth order guess for the order parameter. \begin{figure}[htb] \twofig{fig6a.eps}{fig6b.eps}{6.0} \caption{Both (a) the electrical current throughout an NS junction, and (b) the phase in the superconducting metal, develop naturally from a self-consistent model. The initial order parameter guess $\Delta_0(x)$ assumed zero supercurrent, however a superfluid flow appeared naturally upon reaching self-consistency.} \label{figcurrcons} \end{figure} Fig.~\ref{figcurrcons}(a) shows the electrical current $I(x)$ versus position $x$ inside the superconductor. The numbers beside the lines in Fig.~\ref{figcurrcons}(a) denote the iteration number. For the first iteration ($N=1$), the electrical current dies off within a coherence length of the NS interface, so that the electrical current is not conserved. After the iterative scheme converges ($N=700$) in Fig.~\ref{figcurrcons}(a), we see the electrical current $I(x)$ is constant as a function of position, indicating the electrical current is indeed conserved. In Fig.~\ref{figcurrcons}(b) we plot the order parameter phase inside the superconductor as a function of position. A uniform phase gradient develops inside the superconductor when the iterative scheme has converged to self-consistency, showing that the development of a supercurrent is necessary to guarentee current conservation. We expected this constant order parameter phase gradent in the NS junction, since it is similar to that constant order parameter phase found in the self-consistent solution of the SNS junction.~\cite{rRie1} We can understand these result using our two-fluid picture from section \ref{secwhymore}, and assuming rigidly shifted energy bands. The maximum current of $4e\Delta/h$ is reached when the voltage $eV = \Delta$. At this point the energy bands have shifted up faster than the Fermi level of the normal contact, and thus no direct transmission of electrons across the junction inside the energy range $0<E<eV$ is allowed. Andreev transmission will also be small, as is usually the case in NS junctions. There will therefore be essentially no quasi-particles above the Fermi level of the superconductor, while all states below the Fermi level are filled. The quasi-particle contribution to the current inside the superconductor, $I_{QP}$ in Eq.~(\ref{I_QP}), is essentially zero in the single mode NS junction when $eV < \Delta$. The electrical current is therefore $I = e n v_s$ in this single moded NS junction, the same as for a uniform 1D superconductor. The order parameter magnitude collapses when the superfluid velocity equals the depairing velocity $v_s = v_d$, at a bias voltage $eV = \Delta$. \subsection{Order Parameter Magnitude} \label{opmag} \indent Fig.~\ref{figorderparam}(a) shows the magnitude of the condensation amplitude $F(x)$ as a function of position in both the normal and superconducting metal at zero temperature. The solid line indicates a bias voltage of $V=0$, while the dashed line is for a bias voltage $V = 0.95 V_c$ ($v_s = 0.95 v_d$). The general form of the condensation amplitude $F(x)$ for a ballistic NS junction at equilibrium is well known from earlier non-self-consistent models.~\cite{rMacMill},\cite{rFalk}. We find substantial agreement between these earlier results and our fully self-consistent calculations. In the superconductor, $F(x)$ is suppressed from its bulk value near the NS interface. In the normal metal, $F(x)$ shows behavior quite similar to the low temperature experimental results of Mota~\cite{rMota}. The dotted line in Fig.~\ref{figorderparam}(a) is $F(x) \propto 1/(|x| + x_{0})$, found experimentally by Mota~\cite{rMota}. The value of $x_{0}$ used in Fig.~\ref{figorderparam}(a) is $x_{0} = \xi_{0}$. The fit between the experimental determined form $F(x) \propto 1/(|x| + x_{0})$ and the results of our self-consistent calculation is quite good. The result $F(x) \propto 1/(|x| + x_{0})$ for the pair correlation function at low temperature in an NS junction at equilibrium ($V=0$) was also pointed out by Falk~\cite{rFalk} for the asymptotic limit $x \rightarrow \infty$. \begin{figure}[htb] \twofig{fig7a.eps}{fig7b.eps}{6.0} \caption{The magnitude of the coherence function $F(x)$ (a) changes little when a large flow is present in a NS junction with an applied voltage. The solid line is $F(x)$ when the applied voltage is zero. When the applied voltage is large $.95\Delta$ $F(x)$ shows little change (dashed line) from the junction in equilibrium. [The phase of $F(x)$ changes linearly with postion throughout the NS junction.] (b) In an S'S junction at a temperature above the critical temperature of S', the order parameter is non-zero in the "normal" metal. The finite temperature in combination with a moderate supercurrent causes a supression of the larger gap superconductor. (dashed line)} \label{figorderparam} \end{figure} In an NS junction, the ordering parameter $\Delta(x)$ vanishes in the normal metal because the electron-phonon coupling constant $g(x)$ is zero there. For a finite ordering parameter $\Delta(x)$ to exist inside the normal metal we must have $g(x) \ne 0 $ in the normal metal. One way to achieve a non-zero $g(x)$ in the normal metal is to fabricate an $S^{\prime}S$ junction, where the superconductor $S^{\prime}$ has a smaller critical temperature than S. If we then elevate the temperature so that $T_c > T > T_c^{\prime}$, we effectively form an NS junction where $g(x)$ is not zero inside the normal metal. Fig.~\ref{figorderparam}(b) shows a self-consistent calculation for an such $S^{\prime}S$ junction, where $T_c = 6.6K$, $T_c^{\prime}= 0.66K$, and $T=2K$. Unlike the NS junction, where $\Delta(x)=0$ inside the normal metal, we see a non-zero `tail' of the ordering parameter extending into the normal metal in Fig.~\ref{figorderparam}(b). At zero temperature, the bulk value of the order parameter inside the weaker superconductor is $\Delta^{\prime}(T=0) = 0.1 \Delta (T=0)$. From Fig.~\ref{figorderparam}(b) we see that $\Delta(x=0, T=2K) \simeq 2.5 \Delta^{\prime}(T=0)$, larger than even the bulk value of the order parameter in the weaker superconductor at zero temperature. When a voltage is applied to the $S^{\prime}S$ junction, namely $eV=.7\Delta_{0}$ in Fig.~\ref{figorderparam}(b), the ordering parameter inside $S$ is now suppressed from its bulk value at $T=2K$. This degradation of the order parameter at finite temperature, when the superconductor carries a finite supercurrent, is similar to that of a bulk superconducting wire~\cite{rBag1}. The tail of the ordering parameter extending into $S^{\prime}$ is only slightly changed in the presence of the supercurrent. \subsection{Local Density of States} \label{locdos} \indent In addition to the magnetic susceptibility techniques used by Mota, which explore the condensation amplitude $F(x)$ in the normal metal, another method to experimentally investigating how the ordering parameter $\Delta(x)$ varies near NS interfaces is tunnelling spectroscopy using an STM tip.~\cite{rTess}. We expect that a measurement of the differential conductance $dI/dV$ at the STM tip is proportional to to the local density of states $N(x,E)$. We can calculate the local density of states using the equation \begin{equation} N(x,E) = \frac{1}{\pi}\sum_{p,n} {|u_{p,n}(x,E)|^{2} + |v_{p,n}(x,E)|^{2}}|\frac{dk}{dE}| , \end{equation} where $p$ is the lead index and the quantum number $n = (k,\beta)$. Fig.~\ref{figDOS}(a) shows the local density of states for an NS junction having an applied voltage of $eV=.7\Delta_{0}$. At this bias voltage, the superconductor carries a supercurrent of $v_s = 0.7 v_{d}$. The solid line shows $N(x,E)$ at a position $x = 5 \xi_0$ inside the superconductor. The original peak in the density of states, which occurs at $E = \Delta$ when the superfluid flow $v_s = 0$, splits into two separate peaks. As the energy bands inside the superconductor tilt under the superfluid flow, the band edges move to the energies $E= [1 \pm (v_s/v_{d})]\Delta$, as do the peaks in the density of states. In the normal metal, the density of states is approximately constant for energies of interest. The constant density of states in the normal metal is due to a zero pairing potential $\Delta(x) = 0$ inside the normal metal, since $g(x)=0$ in the normal metal. \begin{figure}[htb] \twofig{fig8a.eps}{fig8b.eps}{6.0} \caption{(a) When a supercurrent is present in an NS junction, the local density of states inside the superconductor shows two peaks. The local density of states inside the normal conductor is constant for all $x < 0$, indicating the lack of any energy gap due to the proximity effect. (b) In an S'S junction when a supercurrent is present, where S' is a superconductor above its transition temperature, structure is seen in the local density of states inside the "normal" S' metal. } \label{figDOS} \end{figure} Fig.~\ref{figDOS}(b) shows the local density of states for an $S^{\prime}S$ junction having $T_c > T > T_c^{\prime}$. We evaluate the local density of states at a position $x = -\xi_{0}$ in the normal metal. The local density of states inside the stronger superconductor is approximately the same as Fig.~\ref{figDOS}(a). For the two different applied voltages, where $v_{s}=0$ (solid) and $v_{s}=.5v_{d}$ (dotted), the presence of the superconductor changes the density of states inside the normal metal. For $v_{s}=0$, there is a depression in the density of states near $E=0$, showing the partial development of an energy gap inside the normal metal. The density of states does not go to zero at $E=0$, since quasi-particles incident from the left contact can still propagate to the position $x = -\xi_{0}$. As the current increases, the density of states inside the normal metal becomes flatter due to injection of quasi-particles from the tilted energy bands inside the superconductor. The two small peaks at $eV=.5\Delta_{0}$ and $eV=1.5\Delta_{0}$ when $v_{s}=.5v_{d}$ are again associated with the tilted energy bands inside S, and are significantly broadened by thermal smearing at $T= 2K$ (which we have ignored in Fig.~\ref{figDOS}(b)). To summarize, while looking to justify our two-fluid model for the effect of superfluid flow in an NS junction, we studied the pair corrlation function $F(x)$ and ordering parameter $\Delta(x)$. We looked at $F(x)$ and $\Delta(x)$ inside both (1) an NS junction at temperature $T=0$, and (2) an S$^{\prime}$S junction at a temperature $T_c > T > T_c^{\prime}$. Here S$^{\prime}$ is a superconductor having an order parameter smaller than $S$, so that $T_c^{\prime} < T_c$. The S$^{\prime}$S junction at this temperature is therefore a type of NS junction. First, the condensation amplitude is $F(x) \simeq x_0 / (|x| + x_0)$, approximately independent of the applied voltage. Second, at the voltage $V_{c} = \Delta_{0}/e$, the ordering parameter of the superconductor collapses, i.e. $\Delta_0 \to 0$. The supercurrent carried inside the superconductor at a voltage $V = V_{c}$ is approximately equal to the Landau depairing current $I_C = e n v_d$. Third, we show how the supercurrent changes the 1-D local density of states per unit energy $N(x,E)$ at various points in the normal and superconducting metals, both for the NS junction and the S$^{\prime}$S junction. The local density of states shows the influence of the superfluid flow. \subsection{Locally (But Not Globally) Gapless Superconductivity} \label{gapless} The phase gradient of $F(x)$ is approximately constant for both the NS junction in Fig.~\ref{figorderparam}(a) and the $S^{\prime}S$ junction in Fig.~\ref{figorderparam}(b). Since the energy gap $\Delta(x)$ decays to zero in Fig.~\ref{figorderparam}(a)-(b), there are regions of local `gapless' superconductivity near the NS interface where $v_s > \Delta(x) / p_F$. However, these regions of local `gapless' superconductivity do not affect bulk properties such as critical current, critical temperature, etc. Since our self-consistent model shows that the supercurrent is approximately constant throughout the superconductor, depairing will occur when $v_s=v_{d}$ everywhere in the superconducting wire. A different type of `global' gapless superconductivity has been discussed by Sanchez and Sols~\cite{rSanSol}. In this proposed gapless superconductor~\cite{rSanSol}, a superconducting wire in contact with an NS junction is postulated to exist when $v_{s} > v_{d}$. It is true that such a non-equilibrium self-consistent solution to the BdG equations with $v_{s} > v_{d}$ does exist for a uniform superconducting wire. However, connecting such a wire in this novel `global' gapless superconducting state to an NS junction imposes the additional constraint that the self-consistency condition in Eq.~(\ref{eCond}) must be satisfied at every point in space. When Eq.~(\ref{eCond}) is satisfied, the magnitude of the order parameter can no longer be a constant in space, as required for the globally gapless solution proposed in Ref.~\cite{rSanSol} to exist. Another way to view the situation proposed in Ref.~\cite{rSanSol} is that, for a given applied voltage, the constraint of current conservation fixes both $\Delta(x)$ and $v_s$, leaving no more degrees of freedom. The order parameter $\Delta(x)$ and superfluid velocity $v_s$ cannot be adjusted independently as required for the bulk solution of Ref.~\cite{rSanSol} to exist in an NS junction. Our self-consistent model shows instead that the ordering parameter of the superconducting wire collapses when the junction voltage is approximately $eV = \Delta$, or equivalently when $v_s = v_d$. In short, we see no possible way to achieve the non-equlibrium conditions necessary for this novel gapless superconducting state of Ref.~\cite{rSanSol} to exist by connecting a superconducting wire to a ballistic NS junction. Fortunately, many other interesting measurements are possible at ballistic NS interfaces without invoking the global gapless superconducting state of Ref.~\cite{rSanSol}. In particular, the excess current larger than the point contact limit, which we find in section \ref{secwhymore}, in no way depends on the gapless state proposed in Ref.~\cite{rSanSol}. \section{Conclusions} \label{conc} \indent It is possible to experimentally observe excess currents larger than the $(4/3)(2e \Delta/h)$ found for ballistic NS point contacts. By varying the width $W_S$ of a superconducting wire in contact with a normal metal having width $W_N$, one can vary the effect of the supercurrent on the energy bands in the superconductor. Varying the widths of the two conductors controls the ratio of the number of conducting modes $\alpha \simeq W_S / W_N$ in the ballistic NS junction. We find the excess current attains a theoretical maximum of $(7/3)(2e \Delta/h)$ when $\alpha \simeq 7/3$. For $1< \alpha < 7/3$ it should be possible to observe discontinuities in the I-V relation of the NS junction when the superfluid velocity $v_{s}$ exceeds the Landau depairing velocity $v_d$. Although these results follow from a simple model which treats all conducting modes as equivalent, we confirmed the qualitative results using a more realistic model which includes the different supercurrent carried in each conducting mode and the scattering between the different modes. These predictions are based on a `two-fluid' solution of the BdG equations, in which current conservation is violated locally near the NS interface. To confirm our that our `two-fluid' type treatment of the superfluid flow in NS junctions generates qualitatively accurate predictions for the I-V relations, we solved the BdG equations self-consistently for a single mode NS junction under an applied bias. Current conservation follows automatically in this self-consistent scheme, and shows that the superfluid velocity is indeed constant throughout the NS junction. The two important features confirmed in this self-consistent solution are (1) the superfluid velocity and terminal currents are the same as required by our `two-fluid' scheme and (2) the order parameter indeed collapses when the superfluid velocity $v_{s}$ equals the depairing velocity $v_d$. We did not perform a completely self-consistent calculation for the multiple moded NS junction, as we believe all the essential elements of this problem (so far as the excess current is concerned) are encompassed in the two-fluid treatment of the multiple-moded NS junction. Having obtained the a self-consistent solution of the BdG equations for an NS junction under bias also enabled us to study the pair correlation function $F(x)$, order parameter $\Delta(x)$, and local density of states $N(x,E)$ in the NS junction. The peak near the superconducting energy gap at $E = \Delta$ in local density of states $N(x,E)$ inside the superconductor is split by the superfluid flow, as can be measured using STM spectroscopy. If the electron-phonon interaction in the normal metal is nearly zero, the density of states in the normal metal is unaffected by the presence of the superconductor. If the normal metal $N$ is a weak superconductor held above its transition temperature, the STM can also measure changes in the local density of states $N(x,E)$ inside the normal metal. Due to the uniform superfluid velocity inside an NS junction, and the reduction of the order parameter near the NS interface, the Landau depairing condition is locally violated and a type of gapless superconductivity occurs locally near the NS interface. \section{Acknowledgments} \indent We thank Supriyo Datta for many useful discussions. We gratefully acknowledge support from the David and Lucile Packard Foundation and from the MRSEC of the National Science Foundation under grant No. DMR-9400415 (PFB). \section{Appendix: Determining the Superfluid Flow Velocity} \label{AppDetFlow} If the pairing potential and electron wavefunctions are determined from a self-consitent solution of the BdG equations, (1) electrical current will be conserved everywhere in space and (2) the superfluid flow will develop naturally as the scheme evolves towards self-consistency (c.f. section \ref{secself1d}). In this appendix we examine a different `globally self-consistent' scheme, where current conservation is only guarenteed at the device leads. In this scheme, the correct value of the superfluid velocity $v_S = \hbar q /m$ is determined by equating the current operator evaluated on the normal side of the NS junction with the same operator evaluated deep (several coherence lengths) inside the superconductor. We then adjust the superfluid velocity (which is a free parameter in the scheme) until the current flowing out of the superconducting lead is the same as the current flow in the normal lead. We consider only a single moded NS junction with $M_N = M_S = 1$, although the results here are easily generalized to multiple conducting modes. The derivation of the electrical current in Ref.~\cite{rDatta1} applies for an NS junction subject to a superfluid flow. Evaluating the current operator inside the normal region, Ref.~\cite{rDatta1} finds the well known formula developed by BTK~\cite{rBTK}, namely \begin{eqnarray} I = (2e/h) \int^{\infty}_{-\infty} & & [1 - T_{Ne,Ne}(E) + T_{Nh,Ne}(E)] \nonumber \\ & & [f(E-eV) - f(E)] dE. \label{I_N} \end{eqnarray} Here we use the notation of Ref.~\cite{rDatta1}, where $T_{Ne,Ne}=R_N$ and $T_{Nh,Ne}=R_A$ are the particle current reflection probabilities for an electron-like quasi-particle incident from the normal metal (right index, `$Ne$') to transmit as a hole or electron in the normal metal (left index). The sum rule from Eq.~(\ref{Pconserv}) then gives $I = I_{QP} + I_A$, with \begin{eqnarray} I_{QP} = (2e/h) \int^{\infty}_{-\infty} & & [T_{Se,Ne}(E) - T_{Sh,Ne}(E)] \nonumber \\ & & [f(E-eV) - f(E)] dE, \label{I_QPTne0} \end{eqnarray} and \begin{eqnarray} I_A = (2e/h) \int^{\infty}_{-\infty} & & [T_{Nh,Ne}(E)+ T_{Sh,Ne}(E) ] \nonumber \\ & & [f(E-eV) - f(E)] dE . \label{2AndTne0} \end{eqnarray} To evaluate the electrical current operator inside the superconductor, we first note that the energy bands inside superconductor subject to a superfluid flow are~\cite{rBag1} \begin{eqnarray} \frac{\hbar^{2}k^{2}}{2m} + \frac{\hbar^{2}q^{2}}{2m} - \mu = \pm \sqrt{(E \mp \hbar^{2} kq/m)^{2} - |\Delta|^{2}} \; . \label{dispersion} \end{eqnarray} The discussion in Ref.~\cite{rBag1} can be extended to show that the particle current incident from the superconductor $J_P = (1/\hbar) (dE/dk)$. Thus, the particle current incident from the superconductor per unit energy is simply $1/h$. Quasi-particles incident from the superconductor will then carry an electrical current of the form \begin{eqnarray} I = \frac{e}{h} \int dE \frac{J_P^{\rm out}}{J_P^{\rm in}} \frac{J_Q^{\rm out}}{J_P^{\rm out}} . \label{exampleJQ} \end{eqnarray} We recognize the particle current transmission coefficient in Eq.~(\ref{exampleJQ}) as \begin{equation} T_{\rm out, in} = \frac{J_P^{\rm out}}{J_P^{\rm in}}. \label{defTP} \end{equation} To evaluate the $J_Q^{\rm out}/J_P^{\rm out}$ term in Eq.~(\ref{exampleJQ}) we use several results from Ref.~\cite{rBag1}. The scattering states inside the superconductor have solutions of the form \begin{eqnarray} \left( \matrix{ u(x) \cr v(x) } \right) = \left( \matrix{ u_{kq} \exp (iqx) \cr v_{kq} \exp (-iqx) } \right) e^{ikx}. \label{uvkq} \end{eqnarray} { \onecolumn The electrical current carried by each occupied state is therefore \begin{equation} J_Q = \left( \frac{\hbar k}{m} \right) \left( |u_{kq}|^2 + |v_{kq}|^2 \right) f + \left( \frac{\hbar q}{m} \right) \left( |u_{kq}|^2 - |v_{kq}|^2 \right) f , \label{defJQ} \end{equation} and the particle current for each occupied state is \begin{equation} J_P = \left( \frac{\hbar k}{m} \right) \left( |u_{kq}|^2 - |v_{kq}|^2 \right) f + \left( \frac{\hbar q}{m} \right) \left( |u_{kq}|^2 + |v_{kq}|^2 \right) f . \label{defJP} \end{equation} The state are normalized so that $|u_{kq}|^2 + |v_{kq}|^2 = 1$. Hence the factor \begin{equation} \frac{J_Q^{\rm out}}{J_P^{\rm out}} = \frac{1}{(|u_{kq}|^2 - |v_{kq}|^2) + (q/k)} + \frac{(q/k)(|u_{kq}|^2 - |v_{kq}|^2)} {(|u_{kq}|^2 - |v_{kq}|^2) + (q/k)}. \label{defJQ/JP} \end{equation} The first term in Eq.~(\ref{defJQ/JP}) one can show is simply the ratio of the density of states in the superconductor to that of the normal metal (for a fixed value of wavevector $k$, not a fixed energy), namely \begin{equation} \tilde{N}^S(E) \equiv \frac{N_S(k)}{N_N(k)} = \frac{1}{(|u_{kq}|^2 - |v_{kq}|^2) + (q/k)} . \label{dosk} \end{equation} We make the translation between wavevector $k$ and energy $E$ inside the superconductor using Eq.~(\ref{dispersion}). The second term in Eq.~(\ref{defJQ/JP}) will be only a minor correction to the first, being nearly zero near the Fermi level, equal to $(q/k) << 1$ over most of the energy range, and equal to 1 only near the bottom of the electron energy bands. The first term in Eq.~(\ref{defJQ/JP}) therefore dominates, being much larger than 1 near the Fermi level and equal to 1 over most of the energy range. Applying the procedure outlined in Ref.~\cite{rDatta1} for construction of the scattering states, and multiplying by their appropriate Fermi occupation factors, we find a total current inside the superconductor of \begin{equation} I_S = I_1 + I_2 . \label{IS} \end{equation} The current $I_1$ arises from first term in Eq.~(\ref{defJQ/JP}) and is \begin{eqnarray} I_1 & = & \frac{2e}{h} \int_{-\infty}^{\infty} [\tilde{N}^S_{Se,out}(E) T_{Se,Ne}(E) - \tilde{N}^S_{Sh,out}(E) T_{Sh,Ne}(E)] [f(E-eV) - f(E)] dE \nonumber \\ & + & \frac{2e}{h} \int_{-\infty}^{\infty} \tilde{N}^S_{Se,out}(E) [2f(E)-1] dE - \frac{2e}{h} \int_{-\infty}^{\infty} \tilde{N}^S_{Se,in}(E) [2f(E)-1] dE . \label{I1s} \end{eqnarray} To obtain Eq.~(\ref{I1s}) we used the sum rule and electron hole symmetry, Eqs.~(C.1) and (C.6) of Ref.~\cite{rDatta1}. We distinguish between $\tilde{N}^S_{Sh,out}(E)$ and $\tilde{N}^S_{Se,in}(E)$ in Eq.~(\ref{I1s}), since the incoming and outgoing electron-like quasi-particles have different densities of states. The second term in Eq.~(\ref{defJQ/JP}) results in a small correction current $I_2$, proportional to $(q/k_F)$. The rather cumbersome expression for $I_2$ is \begin{eqnarray} I_2 & = & \frac{2e}{h} \int_{-\infty}^{\infty} [(q/k)(|u_{kq}|^2 - |v_{kq}|^2)]_{Se,out} \tilde{N}^S_{Se,out}(E) T_{Se,Ne}(E) [f(E-eV) - f(E)] dE \nonumber \\ & - & \frac{2e}{h} \int_{-\infty}^{\infty} [(q/k)(|u_{kq}|^2 - |v_{kq}|^2)]_{Sh,out} \tilde{N}^S_{Sh,out}(E) T_{Sh,Ne}(E)] [f(E-eV) - f(E)] dE \nonumber \\ & + & \frac{2e}{h} \int_{-\infty}^{\infty} [(q/k)(|u_{kq}|^2 - |v_{kq}|^2)]_{Se,out} \tilde{N}^S_{Se,out}(E) [2f(E)-1] dE \nonumber \\ & - & \frac{2e}{h} \int_{-\infty}^{\infty} [(q/k)(|u_{kq}|^2 - |v_{kq}|^2)]_{Se,in} \tilde{N}^S_{Se,in}(E) [2f(E)-1] dE . \label{I2s} \end{eqnarray} We again translate between $k$ and $E$ inside the superconductor using Eq.~(\ref{dispersion}), taking care to assign the appropriate branch of the dispersion curve for incoming or outgoing electron- or hole- like particles. The first two terms in $I_2$ are a small correction to the superfluid flow due to additional quasi-particle injection, while the last two terms are a small correction to the equilibrium superfluid flow. A rigorous treatment guarenteeing global current conservation at any temperature would equate the current $I$ from Eq.~(\ref{I_N}) with the current $I_S$ from Eq.~(\ref{IS}), adjusting the superfluid velocity $q$ until $I = I_S$. } \twocolumn We now analyze the validity of our `two-fluid'' procedure from Section \ref{secwhymore}. We henceforth neglect the current $I_2$ as insignificant compared with $I_1$. At zero temperature, the current operator evaluated inside the superconductor we find from from Eq.~(\ref{I1s}) as \begin{eqnarray} I_1(T=0) & = & \frac{2e}{h} \int_{0}^{eV} [\tilde{N}^S_{Se,out}(E) T_{Se,Ne}(E) \nonumber \\ & - & \tilde{N}^S_{Sh,out}(E) T_{Sh,Ne}(E)] dE \nonumber \\ & + & \frac{4e\Delta}{h} (v_s/v_d). \label{I1sT0} \end{eqnarray} The zero temperature limit of Eq.~(\ref{I_N}) is given in Section \ref{secwhymore} as Eqs.~(\ref{I_QP})-(\ref{2And}). The second term in Eqs.(\ref{I1sT0}) is the superfluid flow term $I_C$. So we can certainly identify $I_A$ from Eq.~(\ref{2And}) with $I_C$. However, Eq.~(\ref{I_QP}) for $I_{QP}$ is not exactly equal to the first term in Eq.~(\ref{I1sT0}), the difference being the additional factor of the superconducting density of states $\tilde{N}^S(E)$ in Eq.~(\ref{I1sT0}). For the Cooper pair flow away from the NS interface (which we are considering in this paper), and for the energy range between $0$ and $eV$, the outgoing hole-like quasi-particle conduction channel opens at a lower energy than the electron-like quasi-particle channel (see Fig.~\ref{figBand}). This means the first term in Eq.~(\ref{I1sT0}) will be larger and more negative than $I_{QP}$ from Eq.~(\ref{2And}), requiring a larger value of the superfluid velocity $v_s$ at each value of the applied voltage $V$ than in the two-fluid model of Section \ref{secwhymore}. We conclude that the treatment in Section \ref{secwhymore} therefore underestimates the effect of superfluid flow on the excess current.
\section{Introduction} Recently, one of us [1] has shown that, under certain physically reasonable conditions, a generic gravitational collapse developing from a regular initial state cannot lead to the formation of a final state resembling the Kerr solution with $a^{2}>m^{2}$---i.e., of a naked singularity accompanied by closed timelike curves. This result supports the validity of Penrose's cosmic censorship hypothesis [2] and suggests that there may exist some deeper connection between cosmic censorship and the chronology protection conjecture put forward by Hawking [3]. An important role in this result plays the so-called {\em inextendibility condition} (see Sec. II), which is assumed to be satisfied for certain incomplete null geodesics. This condition enables one to rule out artificial naked singularities that could easily be created by simply removing points from otherwise well-behaved spacetimes. The inextendibility condition is based on the idea that physically essential singularities should always be associated with large curvature strengths, which are in turn usually associated with the focusing of Jacobi fields along null geodesics. It is easily seen that the inextendibility condition will always hold for null geodesics terminating at the so-called {\em strong curvature singularities} defined by Tipler [4] (see below). Singularities of this type are sometimes considered to be the {\em only} physically reasonable singularities (cf., e.g., [5,6]). However, strong curvature singularities can exist only if the curvature in their neighborhood diverges strong enough [7], while it is not unlikely that some singularities occurring in generic collapse situations will involve a weaker divergence of the curvature. In fact, one cannot {\em a priori} exclude the existence of some ``real'' singularities near which the curvature would remain even bounded (such singularities occur, for example, in Taub-NUT space). Accordingly, since we still have no fully accepted {\em necessary} condition on the behavior of the curvature near generic singularities, one should try to prove any cosmic censorship result under as weak a curvature condition as possible. It would be therefore of interest, in view of the mentioned censorship result [1], to know what are curvature conditions for the occurrence of singularities corresponding to the inextendibility condition. Furthermore, the inextendibility condition has also been used in proving some other recent results [8,9] that restrict a class of possible causality violations in classical general relativity. In this paper, we formulate and prove a theorem that establishes some relations between the inextendibility condition and the rate of growth of the Ricci curvature along incomplete null geodesics. This theorem shows that the inextendibility condition may hold for a much more general class of possible singularities than only those of the strong curvature type. Our theorem will be stated in Sec. II of the paper. In Sec. III we present a proof of the theorem; our main mathematical tool in this proof is a Sturm-type comparison lemma for nonoscillatory solutions of second-order differential equations. In Sec. IV we give a few concluding remarks; in particular, we argue that some earlier cosmic censorship results obtained for strong curvature singularities can be extended to singularities corresponding to the inextendibility condition. \section{The theorem} To begin with, we clearly need to recall the precise formulation of the inextendibility condition. Let $\eta(t)$ be an affinely parametrized null geodesic, and let $Z_{1}$ and $Z_{2}$ be two linearly independent spacelike vorticity-free Jacobi fields along $\eta(t)$. The exterior product of these Jacobi fields defines a spacelike area element, whose magnitude at affine parameter value $t$ we denote by $A(t)$. If we now introduce the function $z(t)$ defined by $A(t)\equiv z^{2}(t)$, then one can show [4] that $z(t)$ satisfies the following equation: \begin{equation} \frac{d^{2}z}{dt^{2}}+\frac{1}{2}(R_{ab}K^{a}K^{b}+ 2\sigma^{2})z=0, \end{equation} where $K^{a}$ is the tangent vector to $\eta(t)$ and $\sigma^{2}$ is a non-negative function of $t$ defined as follows: $2\sigma^{2}\equiv \sigma_{mn}\sigma^{mn}$ $(m,n=1,2)$. Here $\sigma_{mn}$ is the shear tensor (see [10], p. 88) that satisfies the equation [4]: \begin{equation} \frac{d}{dt}\sigma_{mn}=-C_{manb}K^{a}K^{b}- \frac{2}{z}\frac{dz}{dt}\sigma_{mn}. \end{equation} In the following, by $M$ we shall denote a spacetime, i.e., a smooth, boundaryless, connected, four-dimensional Hausdorff manifold with a globally defined $C^{2-}$ Lorentz metric. \vspace{3mm} {\em Definition: Let $\eta: (0,a]\rightarrow M$ be an affinely parametrized, incomplete null geodesic. Assume also that $\eta(t)$ generates an achronal set, i.e., a set such that no two points of it can be joined by a timelike curve. Then $\eta(t)$ is said to satisfy the} {\bf inextendibility condition} {\em if for some affine parameter value $t_{1}\in (0,a)$ there exists a solution $z(t)$ of Eq. (1) along $\eta(t)$ such that $z(t_{1})=0$, $dz/dt|_{t_{1}}\neq 0$ and $\lim_{t\rightarrow 0}z(t)=0$.} \vspace{3mm} The key idea behind the inextendibility condition is based on the fact that any two zeros of any solution of Eq. (1), which is not identically zero along a given null geodesic, correspond to a pair of conjugate points along the geodesic (see [4]). From Proposition 4.5.12 of Ref. [10] it follows that incomplete null geodesics generating achronal sets cannot contain any pairs of conjugate points. One can thus easily show [8] that if a geodesic $\eta: (0,a]\rightarrow M$ satisfies the inextendibility condition, then there is $no$ extension of the spacetime $M$, preserving all the above mentioned properties of $M$, in which $\eta(t)$ could be extended beyond a point $\eta(0)$. This means, according to the standard interpretation, that $\eta(t)$ should then approach a genuine singularity of the spacetime $M$ at affine parameter value 0. [Formally, this singularity has the same status as those predicted by the familiar singularity theorems [10], because these theorems predict in fact the existence of incomplete causal (usually null) geodesics in maximally extended spacetimes satisfying just the same topological and smoothness conditions as those imposed on $M$.] Let us now compare the inextendibility condition with the concept of a strong curvature singularity [4]. Consider a null geodesic $\lambda: (0,a]\rightarrow M$ that terminates in a strong curvature singularity at affine parameter value 0. This means that every solution $z(t)$ of Eq. (1) along $\lambda(t)$, which vanishes for at most finitely many points in $(0,a]$, satisfies $\lim_{t\rightarrow 0}z(t)=0$ (cf. Ref. [5], p. 160). Suppose now that $\lambda(t)$ generates an achronal set; then any solution of Eq. (1), which is not identically zero along $\lambda(t)$, cannot vanish for any two points in $(0,a]$ by the argument with conjugate points mentioned above. Thus, for {\em all} $t_{1}\in (0,a]$ and for {\em all} solutions $z(t)$ of Eq. (1) along $\lambda(t)$ with initial conditions $z(t_{1})=0$ we will have $\lim_{t\rightarrow 0}z(t)=0$. It is thus clear that any null geodesic terminating in Tipler's strong curvature singularity and generating an achronal set must always satisfy the inextendibility condition. Notice also that the terms ``all'' emphasized above imply, via Eqs. (1) and (2), that $\lambda(t)$ can terminate in the strong curvature singularity only if the curvature diverges strong enough along $\lambda(t)$ as $t\rightarrow 0$, while the inextendibility condition could actually be satisfied for $\lambda(t)$ even if the curvature along it would remain bounded. Indeed, the theorem stated below makes it clear [see condition (i)] that the curvature need not necessarily diverge along geodesics satisfying the inextendibility condition. \vspace{3mm} {\bf Theorem:} \, {\em Let $\eta: (0,a]\rightarrow M$ be an affinely parametrized, incomplete null geodesic generating an achronal set. Suppose that the Ricci tensor term $r(t)\equiv R_{ab}K^{a}K^{b}$ along $\eta(t)$, where $t$ is the affine parameter and $K^{a}$ is the tangent vector to $\eta(t)$, obeys at least one of the following conditions.} \\ (i) {\em There exists an affine parameter value $b\in (0,a)$ such that $\inf\{r(t)| \, 0<t<b\}\geq 2(\pi/b)^{2}$.} \\ (ii) {\em There exist an affine parameter value $c\in (0,a)$ and a constant $\mu\in (0,2)$ such that $r(t)\geq \kappa t^{-\mu}$ for all $t\in (0,c]$, where $\kappa = (2/3)(33-26\mu+5\mu^{2})c^{\mu-2}$.} {\em Then $\eta(t)$ satisfies the inextendibility condition. } \vspace{3mm} {\em Remark 1:} From the proof of this theorem, which is given below, it may be seen that the parameter values $b$ and $c$ mentioned above in conditions (i) and (ii) correspond to the parameter value $t_{1}$ occurring in the definition of the inextendibility condition. {\em Remark 2:} Since in the theorem $\eta(t)$ is assumed to be a generator of an achronal set, $\eta(t)$ cannot contain any pair of conjugate points, and so one can expect that there should exist an {\em upper} limit on the rate of growth of the curvature along $\eta(t)$. Indeed, from Theorems (3) and (4) of Ref. [11] it follows immediately that the Ricci tensor term $r(t)$ along $\eta(t)$ must satisfy the following two conditions: (1) there is no affine parameter value $b'\in (0,a]$ such that $\inf\{r(t)|\, 0<t<b'\}>8(\pi/b')^{2}$; and (2) if $r(t)\geq 0$ on $\eta(t)$, then $\lim_{t\rightarrow 0}\inf t^{2}r(t)\leq 1/2$. Similar restrictions on the growth of the Weyl part of the curvature along $\eta(t)$ can be obtained from Proposition 2.2 of Ref. [12]. In the context of our theorem, it is worth recalling the analogous results obtained by Clarke and Kr\'olak [7] for singularities of the strong curvature type. They have been obtained for two definitions of a strong curvature singularity: the original one formulated by Tipler [4] and its modification proposed by Kr\'olak [6]. According to these results, if a null geodesic $\eta: (0,a]\rightarrow M$ terminates at affine parameter value 0 in a strong curvature singularity defined by Tipler (resp., by Kr\'olak), then there must exist some affine parameter value $c\in (0,a]$ such that $R_{ab}K^{a}K^{b}>A t^{-2}$ (resp., $R_{ab}K^{a}K^{b}>A t^{-1}$) on $(0,c]$, where $K^{a}$ is the tangent vector to $\eta(t)$, $t$ is the affine parameter, and $A$ is some fixed positive constant. [Or very similar conditions on the rate of growth of the Weyl part of the curvature along $\eta(t)$ must be satisfied; see Corollary 2 of Ref. 7.] Comparing these results with condition (ii) of our theorem we see that singularities of the strong curvature type involve a considerably stronger divergence of the Ricci tensor term $R_{ab}K^{a}K^{b}$ than singularities corresponding to the inextendibility condition. There may thus exist a large class of curvature singularities that are not strong in the sense of the definition of Tipler or Kr\'olak, but they may still satisfy the inextendibility condition. Note also that the above conditions for strong curvature singularities are the {\em necessary} ones, whereas conditions (i) and (ii) of our theorem are only {\em sufficient} to ensure that the inextendibility condition does hold for a given geodesic. This implies that the inextendibility condition might be satisfied in more general situations than only those characterized by conditions (i) and (ii). \section{Proof of the theorem} \vspace{4mm} Now we shall prove the theorem; our main tool in this proof will be the following comparison lemma. \vspace{3mm} {\em Lemma (The comparison lemma): Suppose that $u(s)$ is a solution of the equation \begin{displaymath} \frac{d^{2}u}{ds^{2}}+F(s)u(s)=0 \end{displaymath} on an interval $(a,b]$ with initial conditions: $u(b)=0$ and $du/ds|_{b}\neq 0$. Let $v(s)$ be a solution of \begin{displaymath} \frac{d^{2}v}{ds^{2}}+G(s)v(s)=0 \end{displaymath} on $(a,b]$ such that $v(b)=0$, $dv/ds|_{b}=du/ds|_{b}$ and $v(s)>0$ on $(a,b)$. Assume also that $F(s)$ and $G(s)$ are piecewise continuous on $(a,b]$, and let $G(s)\geq F(s)$ on $(a,b]$. Then $u(s)\geq v(s)$ on $(a,b]$. } \vspace{3mm} $Proof:$ \, The proof of this lemma is based essentially on Theorem 1.2 of Ref. [13], p. 210. To apply this theorem in its original form, it is convenient to reparametrize both of the equations in the lemma introducing the parameter $t=-s$ instead of $s$. Note that this reparametrization does not change the form of the equations. Clearly, we shall now have established the lemma if we show that for any $c\in (a,b)$, $u(t)\geq v(t)$ on $[-b,-c]$. Consider the ratio $u(t)/v(t)$. Since $v(t)>0$ on $(-b,-a)$, it is well defined on $(-b,-c]$. Using l'Hospital's rule, we get \begin{displaymath} \lim\limits_{t\rightarrow -b}\frac{u(t)}{v(t)}=1. \end{displaymath} Therefore, as $v(t)>0$ on $(-b,-c]$, to show that $u(t)\geq v(t)$ on $[-b,-c]$, it suffices to show that \begin{displaymath} \frac{d}{dt}\left[\frac{u(t)}{v(t)}\right]\geq 0 \end{displaymath} on $(-b,-c]$. It is easy to see that this inequality holds if \begin{equation} \frac{v(t)}{\dot{v}(t)}\geq \frac{u(t)}{\dot{u}(t)} \end{equation} on $(-b,-c]$, where the overdot denotes the first derivative with respect to $t$. Since $F(t)$ and $G(t)$ are piecewise continuous on $[-b,-c]$, by Theorem 1.2 of Ref. 13, p. 210, we have \begin{displaymath} \tan^{-1}\left[\frac{v(t)}{\dot{v}(t)}\right]\geq \tan^{-1}\left[\frac{u(t)}{\dot{u}(t)} \right] \end{displaymath} for all $t\in [-b,-c]$. Thus, as $\tan^{-1}$ is an increasing function, the inequality (3) does hold as it is desirable. \hfill $\Box$ {\em Proof of the theorem:} (Part I) Suppose the condition (i) is satisfied. Let $z_{0}(t)$ be a solution of Eq. (1) along $\eta(t)$ such that $z_{0}(t)$ is not identically zero on $(0,b]$ and $z_{0}(b)=0$, where $b$ is the parameter value mentioned in condition (i). Clearly, such a solution will always exist. Since $\eta(t)$ generates an achronal set, $z_{0}(t)$ can vanish nowhere in $(0,b)$; otherwise $\eta(t)$ would have a pair of conjugate points in $(0,b]$ (see Ref. [4]), which would contradict, by Proposition 4.5.12 of Ref. [10], the achronality of $\eta(t)$. Notice also that Eq. (1) is linear, and so the function $-z_{0}(t)$ will be a solution of Eq. (1) as well. Thus, as $z_{0}(t) \neq 0$ on $(0,b)$, without loss of generality we can assume that $z_{0}(t)>0$ on $(0,b)$. This implies, as $z_{0}(b)=0$, that $dz_{0}/dt|_{b}\leq 0$. Since $z_{0}(t)>0$ on $(0,b)$, and condition (i) holds, from Eq. (1) we see at once that $z_{0}(t)$ must be a concave function on $(0,b]$. This makes it obvious that $dz_{0}/dt|_{b}\neq 0$, and so we must have $dz_{0}/dt|_{b}=\alpha<0$. Let us now define the function $z_{1}(t)\equiv-(1/\alpha)z_{0}(t)$. As Eq. (1) is linear, it is clear that $z_{1}(t)$ will be a solution of Eq. (1) along $\eta(t)$; notice also that $z_{1}(t)>0$ on $(0,b)$, $z_{1}(b)=0$ and $dz_{1}/dt|_{b}=-1$. Consider now the equation \begin{equation} \frac{d^{2}x}{dt^{2}}+\omega x(t)=0, \end{equation} where $\omega=\frac{1}{2}\inf\{r(t)|\, 0<t\leq b\}$ and $r(t)$ is the function defined in the theorem. Notice that $\omega>0$ by condition (i). Let $x_{1}(t)$ be a solution of Eq. (4) on $(0,b]$ with initial conditions: $x_{1}(b)=0$ and $dx_{1}/dt|_{b}=-1$. It is a simple matter to see that $x_{1}(t)=\omega^{-1/2}\sin[\omega^{1/2} (b-t)]$. Let us now apply the comparison lemma to the equations (1) and (4) and their solutions $z_{1}(t)$ and $x_{1}(t)$. Since $\omega\leq\frac{1}{2}r(t)$ on $(0,b]$, by the comparison lemma we must have $x_{1}(t)\geq z_{1}(t)$ on $(0,b]$. Consequently, as $z_{1}(t)>0$ on $(0,b)$, we obtain $x_{1}(t)>0$ on $(0,b)$. This implies, by the above form of $x_{1}(t)$, that $\omega\leq(\pi/b)^{2}$. But $\omega\geq (\pi/b)^{2}$ by condition (1). We must thus have $\omega=(\pi/b)^{2}$, which gives $\lim_{t\rightarrow 0}x_{1}(t)=0$. Therefore $\lim_{t\rightarrow 0}z_{1}(t)=0$ since $x_{1}(t)\geq z_{1}(t)> 0$ on $(0,b)$. This means that $\eta(t)$ does satisfy the inextendibility condition. (Part II) The task is now to prove the theorem in the case when condition (ii) holds. For this purpose, let us consider the following equation \begin{equation} \frac{d^{2}y}{dt^{2}}+Bt^{-\mu}y(t)=0 \end{equation} on $(0,c]$, where $B=\kappa /2$, and $\kappa$, $\mu$ and $c$ are some fixed constants mentioned in the condition (ii). Let $y_{1}(t)$ be a solution of this equation with initial conditions: $y_{1}(c)=0$ and $dy_{1}/dt|_{c}=-1$. Let $z_{2}(t)$ be a solution of Eq. (1) along $\eta(t)$ such that $z_{2}(c)=0$ and $dz_{2}/dt|_{c}=-1$. [There is no loss of generality in assuming $z_{2}(t)$ to exist; the existence of $z_{2}(t)$ can be established in the same manner as the existence of the analogous solution $z_{1}(t)$ considered in the first part of the proof.] Clearly, the solution $z_{2}(t)$, just as $z_{1}(t)$, can vanish nowhere in $(0,c)$ by the argument with conjugate points. Therefore, as $dz_{2}/dt|_{c}=-1$, we must have $z_{2}(t)>0$ on $(0,c)$. Let us now apply the comparison lemma to the equations (1) and (5) and their solutions $z_{2}(t)$ and $y_{1}(t)$. By condition (ii) we have $r(t) \geq \kappa t^{-\mu}$ on $(0,c]$. Thus by the comparison lemma, we must have $y_{1}(t)\geq z_{2}(t)$ on $(0,c]$. Of course, in order to prove the theorem, it suffices to show that $\lim_{t\rightarrow 0}z_{2}(t)=0$. Thus, as $y_{1}(t)\geq z_{2}(t)>0$ on $(0,c)$, to complete the proof it suffices to show that $\lim_{t\rightarrow 0}y_{1}(t)=0$. We shall show below that $y_{1}(t)$ does possess this property. To this end, let us first find the general solution of Eq. (5). It is easy to check that if one puts $x=t$, $\alpha=1/2$, $\beta= 2\sqrt{B}(2-\mu)^{-1}$, $\gamma=(2-\mu)/2$ and $n=(2-\mu)^{-1}$ into the equation (4.1) of Ref. [14], p. 138, then this equation reduces to our equation (5). Thus, according to the solution (4.3) of Eq. (4.1) of Ref. [14], our equation (5) has the following general solution \begin{equation} y(t)=t^{1/2}[C_{1}J_{n}(\beta t^{\gamma})+C_{2}Y_{n}(\beta t^{\gamma})], \end{equation} where $C_{1}$ and $C_{2}$ are arbitrary constants of integration, and $J_{n}(\beta t^{\gamma})$ and $Y_{n}(\beta t^{\gamma})$ are the Bessel functions of order $n$, of the first and second kind, respectively. Since $\mu\in (0,2)$, from the above relations it follows that $1/2<n<\infty$, $\sqrt{B}<\beta<\infty$ and $0<\gamma<1$. Let us recall that any Bessel function of the first kind has infinitely many positive zeros (cf., e.g., [15], p. 29). Let $j_{n,1}$ be the first positive zero of the function $J_{n}(\beta t^{\gamma})$, i.e., $J_{n}(j_{n,1})=0$ and $J_{n}(\beta t^{\gamma})\neq 0$ as long as $0<\beta t^{\gamma}<j_{n,1}$. Since $n>1/2$, $j_{n,1}$ must satisfy the following relation (see Eq. (2) of Ref. [15], p. 29): \begin{equation} j_{n,1}<2[(n+1)(n+5)/3]^{1/2}. \end{equation} For $J_{n}(\beta t^{\gamma})$ we now define $L$ to be the number such that $j_{n,1}=L\beta c^{\gamma}$. Putting this into (7), and taking into account the fact that $\beta =(2\kappa)^{1/2}(2-\mu)^{-1}$, $\kappa=3^{-1}(66-52\mu+10\mu^{2})c^{\mu-2}$, $\gamma=(2-\mu)/2$ and $n=(2-\mu)^{-1}$, we readily find that $L^{2}<1$. Consider now equation (5) with $B$ replaced by $B'=L^{2} B$. Let $y_{2}(t)$ be a solution of this equation on $(0,c]$ with initial conditions: $y_{2}(c)=0$ and $dy_{2}/dt|_{c}=-1$. The general form of this solution is given by (6), where $\beta$ should be replaced by $\beta'= 2\sqrt{B'}(2-\mu)^{-1}$ (notice that $\beta'= L\beta$). Let us now insert the initial conditions for $y_{2}(t)$ into this general solution in order to determine for $y_{2}(t)$ the constants $C_{1}$ and $C_{2}$ occurring in (6). To find the first derivative of the general solution (6), we use the following recurrence formula \begin{displaymath} \frac{dJ_{n}(x)}{dx}=-J_{n+1}(x)+\frac{n}{x}J_{n}(x), \end{displaymath} which is also valid for $Y_{n}(x)$ (see [15], p. 197). We can now easily calculate the constants $C_{1}$ and $C_{2}$; the result is as follows \begin{equation} C_{1}=\frac{Y_{n}(\beta'c^{\gamma})}{\beta'\gamma c^{\gamma -1/2}\left[Y_{n}(\beta'c^{\gamma}) J_{n+1}(\beta'c^{\gamma})-Y_{n+1}(\beta'c^{\gamma}) J_{n}(\beta'c^{\gamma})\right]} \end{equation} and \begin{equation} C_{2}=\frac{-J_{n}(\beta'c^{\gamma})}{\beta'\gamma c^{\gamma -1/2}\left[Y_{n}(\beta' c^{\gamma}) J_{n+1}(\beta' c^{\gamma})-Y_{n+1}(\beta' c^{\gamma})J_{n}(\beta' c^{\gamma})\right]}. \end{equation} As $\beta'=L \beta$, from the above definition of $L$ it is clear that $\beta' c^{\gamma}=j_{n,1}$. Thus $J_{n}(\beta' c^{\gamma}) =0$ and the numerator in (9) must vanish. As $J_{n}(\beta' c^{\gamma})=0$, the denominator in (9) can vanish only if $Y_{n}(\beta'c^{\gamma}) J_{n+1}(\beta' c^{\gamma})=0$. But the Bessel functions $J_{n+1}$ and $Y_{n}$ cannot have any common zeros with the Bessel function $J_{n}$ (see [15], pp. 29-32), and so the denominator in (9) cannot vanish. We thus have $C_{2}=0$ and, by (6) and (8), the solution $y_{2}(t)$ can be written as follows \begin{equation} y_{2}(t)=C_{1}t^{1/2}J_{n}(\beta' t^{\gamma}), \end{equation} where $C_{1}=[\beta'\gamma c^{\gamma -1/2}J_{n+1}(\beta' c^{\gamma})]^{-1}$. Let us now compare the solutions $y_{1}(t)$ and $y_{2}(t)$ by means of the comparison lemma. Recall that $y_{1}(t)$ is a solution of equation (5) with $B=\kappa/2$, while $y_{2}(t)$ is a solution of the same equation with $B$ replaced by $B'=L^{2}\kappa/2$. Since $L^{2}<1$, by the comparison lemma we must have $y_{2}(t)\geq y_{1}(t)$ for all $t\in (0,c]$. We recall that any Bessel function $J_{k}(x)$ of the first kind with real $x$ and $k>0$ is continuous at $x=0$ (cf. [15], p. 182). Thus, as $n>1/2$ and $0<\gamma<1$, from (10) it follows immediately that $\lim_{t\rightarrow 0}y_{2}(t)=0$. Therefore, as $y_{2}(t)\geq y_{1}(t)>0$ on $(0,c)$, we obtain $\lim_{t\rightarrow 0}y_{1}(t)=0$, which completes the proof. \hfill $\Box$ \section{Concluding remarks} We have been concerned in this paper with the problem of determining what are curvature conditions for the occurrence of singularities corresponding to the inextendibility condition. We have found two such sufficient conditions concerning the behavior of the Ricci tensor term $R_{ab}K^{a}K^{b}$ along incomplete null geodesics---these are conditions (i) and (ii) of the theorem stated in Sec. II. This theorem shows that the inextendibility condition may hold for a considerably larger class of possible singularities than only those of the strong curvature type. In particular, condition (i) of the theorem shows that the inextendibility condition may hold even if the curvature along incomplete geodesics would remain bounded. In this context, it is worth recalling that singularities predicted by the famous singularity theorems [10] can be interpreted as regions of the universe at which the normal classical spacetime picture and/or certain energy conditions break down, and this may occur in regions where the curvature, though extremely large, still remains finite. Accordingly, if one attempts to establish, for example, whether or not these singular regions will conform to any cosmic censorship principle, it would be well to try to characterize, if necessary, incomplete geodesics terminating in these regions by a condition that may hold even if the curvature along the geodesics would remain bounded. One possible candidate for such a condition may thus be the inextendibility condition. It should also be stressed here that some earlier cosmic censorship theorems [6,16,17] proved for strong curvature singularities can be extended to singularities corresponding to the inextendibility condition. To see this, let us first recall that these theorems show, briefly, that under certain restrictions imposed on the causal structure, strong curvature singularities are censored (see Refs. [6,16,17] for details). Proofs of these theorems are, in essence, alike. In a brief outline, they run as follows. First, one shows that if the theorem under consideration were false, then there would have to exist a sequence $\{\mu_{i}\}$ of future endless, future complete null geodesics converging to a null geodesic $\mu$ that terminates in the future at a strong curvature singularity. One also shows that $\mu$ and all the $\mu_{i}$ must be generators of achronal sets. As all $\mu_{i}$ are achronal, none of them can have a pair of conjugate points, and so any irrotational congruence of Jacobi fields along any $\mu_{i}$ cannot be refocused. As $\{\mu_{i}\}$ converges to $\mu$, this must then imply, by continuity, that any irrotational congruence of Jacobi fields along $\mu$ cannot be refocused as well. However, as $\mu$ terminates in a strong curvature singularity, all irrotational congruences of Jacobi fields along $\mu$ should be refocused. This gives the required contradiction. It is not difficult to see, however, that this contradiction can equally well be obtained if $\mu$ would be assumed to satisfy the inextendibility condition, for this condition holds if at least one irrotational congruence of Jacobi fields along a given geodesic is refocused. It is thus clear that the censorship theorems given in Refs. [6,16,17] are unnecessarily restricted to strong curvature singularities and they can be extended to singularities corresponding to the inextendibility condition. \acknowledgments This research was supported in part by the Polish State Committee for Scientific Research (KBN) under Grant No. 2 P03B 073 15.
\section{Introduction} The possible changes of space-time properties at small (Plank) scale now extensively discussed $\cite{Aha,Dop}$. Due to the absence of any experimental information it seems instructive to look for some directions exploring attentively the standard Quantum Physics space-time structure. Some years ago Aharonov and Kaufherr have shown that in nonrelativistic Quantum Mechanics (QM) the correct definition of physical reference frame (RF) must differ from commonly accepted one, which in fact was transferred copiously from Classical Physics $\cite{Aha}$. The main reason is that to perform exact quantum description one should account the quantum properties not only of studied object, but also RF, despite the possible practical smallness. The most simple of this RF properties is the existence of Schroedinger wave packet of free macroscopic object with which RF is usually associated $\cite {Schiff}$. Then it introduces additional uncertainty into the measurement of object space coordinate in this RF. Furthermore this effect results in the states transformations between two such RFs which includes quantum corrections to the standard Galilean group transfromations $\cite{Aha}$. Algebraic and group theorettical structure of this transformation was studied in $\cite{Tol}$. In their work Aharonov and Kaufherr formulated Quantum Equivalence Principle (QEP) in nonrelativistic QM - all the laws of Physics are invariant under transformations between both classic and this finite mass RFs which called quantum RFs. The importance of RF quantum properties account was shown already in Quantum Gravity and Cosmology studies $\cite {Rov,Unr,Dew}$ and will be considered here in connection with the time problem in quantum gravity. In this paper the consistent relativistic covariant theory of quantum RFs formulated, our preliminary results were published \cite{May2}. In this theory no new {\it ad hoc} hypothesis introduced; all calculations are performed in the standard QM formalism. It will be shown that the transformation of the particle state between two quantum RFs obeys to relativistic invariance principles, but differs from standard Poincare Group transformations, due to quantum relativistic correction for RF motion. Solving the evolution equation for quantum clocks models the proper time in moving quantum RF calculated and the related effects of RF momentum quantum fluctuations revealed. This clocks model applied for the analysis of the space-time structure of canonical quantum gravity $\cite {Rov}$. In chap.2 canonical formalism for quantum RFs described. In chap.3 we study quantum clocks models and obtain relativistic proper time for quantum RF. In chap.4 the relativistic evolution equations and unitary transformations for quantum RFs described. We'll consider also RF quantum motion in gravitation field where gravitational 'red shift' results in additional clocks time fluctuations. \section{Quantum Coordinates Transformations } For the beginning we'll consider Quantum Measurements problems related to Quantum RFs model. In QM framework the system defined as RF should be able to measure the observables of studied quantum states and so include the measuring device - detector D. As the realistic example we can regard the photoemulsion plate or the diamond crystal which can measure the particle position and simultaneously record it. Despite the multiple proposals up to now the established theory of collapse doesn't exist $\cite{May,Desp}$. Yet our problem premises doesn't connected directly with any state vector collapse mechanism and and it's enough to detailize standard QM collapse postulate of von Neumann measurement theory $\cite {Desp}$. We consider RF which consists of finite number of atoms (usually rigidly connected) and have the finite mass. It's well known that the solution of Schroedinger equation for any free quantum system can be factorized as : \begin {equation} \Psi(t)= \sum c_l\Phi^c_l(\vec{R}_c,t)*\phi_l(u_{k},t) \label {A1} \end {equation} where center of mass coordinate $\vec{R}_c=\sum m_i*\vec{r}_i/M$, $c_l$ are the partial amplitudes. $u_k$ describes the internal degrees of freedom, which for potential forces are reduced to $\vec{r}_{i,j}=\vec{r}_i-\vec{r}_j$ $\cite{Schiff}$. Here $\Phi^c_l$ describes the c.m. motion of the system. It means that the evolution of the system is separated into the external evolution of pointlike particle M and the internal evolution defined by $\phi_l(u_{k},t)$. So the internal evolution is independent of whether the system is localized in some 'absolute' reference frame (ARF) or not. Quantum Field Theory evidences that the factorization of c.m. and relative motion holds true even for nonpotential forces and variable $N$ in the secondarily quantized systems $\cite{Schw}$. Moreover this factorization expected to be correct for nonrelativistic systems where binding energy is much less then its mass $M$, which is characteristic for the real detectors and clocks. Consequently it's reasonable to assume that this factorization fulfilled also for the detector states despite we don't know their exact structure. For our problem it's enough to assume that eq.(\ref{A1}) holds for RF state only in the time interval $T$ from RF preparation moment $t_0$ until the act of measurement starts , i.e. the measured particle $n$ wave packet $\psi_n$ impacts with D. If this factorization holds the space coordinate measured in this RF depends not only on $\psi_n$ but also on $\Phi_l^c$ which permit in principle to study quantum RF effects. In this case the possible factorization violation at later time when the particle state collapse occured is unimportant for us. We regard in our model that all measurements are performed on the quantum pairs ensemble of particles $G^2$ and $F^1$. It means that each event is resulted from the interaction between the 'fresh' RF and particle ,prepared both in the specified quantum states, alike the particle alone in the standard experiment. To illustrate the meaning of Quantum RF consider gedankenexperiment where in ARF the wave packet of RF $F^1$ described by $\psi_1=\eta_1(x)\xi_1(y)\zeta_1(z)$ at time moment $t_0$. The test particle $n$ with mass $m_n$ belongs to narrow beam which average velocity is orthogonal to $x$ axe and its wave function at $t_0$ is $\psi_n=\eta_n(x)\xi_n(y)\zeta_n(z)$. Before they start to interact this system wave function is the product of $F^1$ and $n$ packets. We want to find $n$ wave function for the observer in $F^1$ rest frame. In general it can be done by means of the canonical transformations described below, but in the simplest case when $n$ beam is localized so that $\psi_n$ can be approximated by delta-function $\delta(x-x_b)\delta(y-y_b)\delta(z-z_b)$ n wave function in $F^1$ easily calculated $\psi'_n(\vec{r}'_n)=\eta_1(x'_n-x_b)\xi_1(y'_n-y_b)\zeta_1(z'_n-z_b)$. It shows that if for example $F^1$ wave packet along $x$ axe have average width $\sigma_x$ then from the 'point of view' of observer in $F^1$ each object localized in ARF acquires wave packet of the same width $\sigma_x$ in $F^1$ and any measurement in $F^1$ and ARF will confirm this conclusion. The generalized Jacoby canonical formalism will be applied in our model alternatively to Quantum Potentials used in $\cite{Aha}$. Consider the system $S_N$ of $N$ objects $W^k$ which include $N_f$ frames $F^i$ which have also some internal degrees of freedom and $N_g=N-N_f$ pointlike 'particles' $G^i$. At this stage we can regard both of them as equivalent objects in the relation to their c.m. motion. We'll assume for the beginning that particles and RFs canonical operators $\vec{r}_i,\vec{p}_i$ are defined in absolute (classical) ARF - $F^0$ having very large mass $m_0$, but later this assumption can be abandoned. We'll start with Jacoby canonical coordinates $\vec{u}_j^l$ associated with $F^l$ rest frame, which for $l=1$ equal : \begin{eqnarray} \vec{u}^1_{i}=\frac{\sum\limits^N_{j=i+1}m_j\vec{r}_{j}}{M_{i+1}} -\vec{r}^l_{i} ;\quad 1\le i<N ; \quad \vec{u}_N=\vec{u}_s=\vec{R}_{c} \quad \label{B1} \end{eqnarray} where $M_i=\sum\limits^{N}_{j=i}m_j$. $\vec{u}^l_i$ can be obtained and is the linear combination of $\vec{u}^1_i$. Conjugated to $\vec{u}^l_i$ canonical momentums are : \begin {equation} \vec{\pi}^1_i= \mu_i(\frac{\vec{p}^s_{i+1}}{M_{i+1}}-\frac{\vec{p}_{i}}{m_{i}}) ,\quad \vec{\pi}_N=\vec{p}_s=\vec{p}^s_1 \label{B2} \end {equation} where $\vec{p}^s_{i}=\sum\limits^{N}_{j=i}\vec{p}_j$ ,and reduced mass $ \mu^{-1}_i=M_{i+1}^{-1}+m_{i}^{-1} $ . The relative coordinates $\vec{r}_j-\vec{r}_1$ can be represented as the linear sum of $\vec{u}^1_i$. They don't constitute canonical set due to the quantum motion of $F^1$ $\cite{Aha}$. The Hamiltonian of $S_N$ motion in ARF is expressed also via momentums $\vec{\pi}^1_i$ : \begin {equation} \hat{H}=\sum\limits_{i=1}^{N}\frac{\vec{p}_i^2}{2m_i}= \frac{\vec{p}_s^2}{2M}+ \sum\limits_{j=1}^{N-1}\frac{(\vec{\pi}^{1}_j)^2}{2\mu_j} =\hat{H}_s+\hat{H}_c \label {B5} \end {equation} In $F^1$ rest frame the true observables are $\vec{\pi}^1_i , \vec{u}^1_i$ and it's impossible to measure $S_N$ observables $\vec{p}_s$ and $\vec{R}_c$. The true Hamiltonian of $S_N$ in $F^1$ should depend on the true observables only , so we can regard $\hat{H}_c$ as the real candidate for its role. It results into modified Schroedinger equation which depends not only of particles masses ,but on observer mass $m_1$ also. Now we'll regard here the alternatve form of this formalism which use Jacoby frame condition (JFC) and is more convenient for the relativistic problem. For the described system $S_N$ Langrangian in ARF $L=\sum\frac{m_i\dot{\vec{r}}_i^2}{2}$ gives $H$ of ($\ref{B5}$) after Legandre transform. If one wish to include ARF motion in this formalism the simplest way is to define formally $L'=L+\frac{m_0\dot{\vec{r}}_0}{2}$. It gives $N+1$ canonical momentums : $\vec{p}_j=\frac{\partial L'}{\partial\dot{\vec{r}}_j}$. The new Langrangian $L'$ is formally symmetric relative to the frame choice and it gives the Hamiltonian $H'=H+\frac{\vec{p}_0^2}{2m_0}$ for $H$ of ($\ref{B5}$). Due to it to anchor this momentums and $H'$ to $F^i$ rest frame in which they acquire some values one must broke $L'$ symmetry introducing the frame condition (FC) or kinematical (holonomial) constraint $\cite{Git}$. For ARF rest frame we choose FC $\vec{p}^{\,2n}_0 \approx 0$, where from the formal reasons $n=2$. It means that $\dot{\vec{r}}_0=0$ - RF is at rest relative to itself which seems quite natural, yet it differs from FC used in $\cite{Aha}$. All Classical and QM results are reproduced in this scheme if ARF mass is taken infinite. $S_N$ quantization in $F^1$ performed with Hamitonian $\hat{H'}$ and FC regarded as the operator which obeys to Dirack rules for the first order constraints $\cite {Dir,Git}$. Galilean-like passive transformations from ARF to $F^1$ and back can be found introducing FC also for $F^1$ $\vec{p}^{\,2n}_{11}\approx 0$, where $\vec{p}_{1i}$ are the canonical momentums in $F^1$. $S_{N+1}$ unitary transformation from ARF to $F^1$ is convenient to write via the $F^{0,1}$ total momentum $\vec{p}_f=\vec{p}_0+\vec{p}_1$ and $F^0,F^1$ relative momentum $\vec{\pi}_f$ conserving other momentums $\vec{p}_i$. Their conjugated coordinates $\vec{r}_f,\vec{u}_f$ have the standard form of ($\ref {B1}$). In this notations the transformation from $F^0$ to $F^1$ is equal to : \begin {equation} U_{1,0}=P_f e^{i a_f\vec{p}_f\vec{r}_f}e^{-i\vec{p}_f\vec{b}_s} \prod_{i=2}^{N}e^{-im_i\vec{r}_i\vec{\beta}} \label{BBB} \end {equation} where $a_f=\ln\frac{{m}_0}{m_1}, \vec{b}_s=\frac{M}{m_0}\langle\vec{R}_c\rangle$. $P_f$ is $\vec{r}_f$ reflection (parity) operator. $\vec{\beta}=\frac{\vec{p}_f}{m_1}$ is the operator corresponding to the velocity parameter in Galilean transformation. Under this transformation $\vec{p}_f$ transformed to $\vec{p}_{1f}=\vec{p}_{10}+\vec{p}_{11}$ and $\vec{\pi}_{1f}= \vec{\pi}_{f}$. Alike the transformation from $\vec{p}_j$ to $\vec{\pi}^1_i$ obtained operator $U_{1,0}$ includes the dilatation transformation $\cite{Bar}$. For $N=2$ one obtains $F^1$ momentums and coordinates : \begin{eqnarray} \vec{p}_{10}=(1-\frac{m_0}{m_1})\vec{p}_0-\frac{m_0}{m_1}\vec{p}_1 \quad ;\quad \vec{r}_{10}=-\frac{m_1\vec{r}_1+m_2\vec{r}_2}{m_0}+\vec{b}_s \nonumber \\ \vec{p}_{11}=-\vec{p}_0 \quad ;\quad \vec{r}_{11}=-\vec{r}_0+(1-\frac{m_1}{m_0})\vec{r}_1-\frac{m_2}{m_0}\vec{r}_2 +\vec{b}_s \label {BA} \\ \vec{p}_{12}=-\frac{m_2}{m_1}(\vec{p}_0+\vec{p}_1)+\vec{p}_2 \quad ; \quad \vec{r}_{12}=\vec{r}_2 \nonumber \end{eqnarray} Results for $N>2$ can be easily deduced from this formulaes. It's easy to see that ARF FC transformed into $F^1$ FC. All $\vec{\pi}^1_i$ are conserved and space shift on $\vec{b}_s$ conserves all the distances $ \vec{r}_i-\vec{r}_j$. In the limit where heavy $F^1$ moves nearly classically $U_{01}$ becomes the Galilean momentum transformation with the velocity $\langle\vec{\beta}\rangle$. $S_{N+1}$ Hamitonian in $F^1$ also can be rewritten via new relative momentums $\vec{\pi}_{1j}$ which can be easily derived following ($\ref {B2}$) : \begin{equation} \hat{H}^1=\sum^N_{i=0}\frac{\vec{p}^2_{1i}}{2m_i}=H^1_s+H^1_c =\frac{\vec{p}_{1s}^2}{2M_{N+1}}+ \sum_{j=2}^{N+1}\frac{\vec{\pi}_{1j}^2}{2\mu_{1j}} \label {BYY} \end{equation} The term $\hat{H}^1_s$ describes $S_N$ c.m. motion relative to $F^1$ which doesn't influence on the evolution of $S_{N+1}$ true observables $\vec{\pi}_{1i} , \vec{u}_{1i}$ or $\vec{r}_i-\vec{r}_j$. $\vec{r}_{1i},\vec{p}_{1i}$ aren't $S_{N+1}$ observables for $F^1$ observer, yet $\vec{p}_{1i}$ expectation values can be found from $\vec{\pi}^1_i$ measurements Now we have quantum system $S_{N+1}$ which include ARF and in ARF rest frame we can ascribe to it without any contradictions with QM the state vector which for $N=2$ is equal : $\psi_s(\vec{p}_0,\vec{p}_1,\vec{p}_2)=\varphi(\vec{p}_1,\vec{p}_2) |\vec{p}_0=0\rangle |\vec{p}_1\rangle |\vec{p}_2\rangle$. After $U_{1,0}$ transformation it acquires the similar form in $F^1$ rest frame with $|\vec{p}_{11}=0\rangle$. As the result of this transform we obtain the new canonical coordinates referred to finite mass $F^1$ rest frame. They permit to factorize internal $S_N$ motion and ARF motion and dropping ARF term in $H^1$ of ($\ref{BYY}$) we obtain $S_N$ Hamitonian. Remind that active transformation shifts $G^2$ state $\psi_2$ on the distance $\vec{a}$ and velocity $\vec{\beta}$ relative to RF. Passive $G^2$ transformation means the transition from one RF to another, but for quantum RF with state $\psi_s$ it can't be described by any state shift on $\vec{a},\vec{\beta}$ and have more complicated form. $U_{1,0}$ is such passive transformation and active $G^2$ transformation is the standard Galilean one even in $F^1$ $\cite {Schw}$. In general the quantum transformations in 2 or 3 dimensions should also take into account the possible rotation of quantum RF axes relative to ARF, which introduce additional angular uncertainty into objects coordinates. Thus after performing coordinate transformation $\hat{U}_{A,1}$ from ARF to $F^1$ c.m. we must rotate all the objects (including ARF) around it on the uncertain polar and azimuthal angles $,\phi_1,\theta_1$ which are $F^1$ internal degrees of freedom. We can imagine $F^1$ axes as some solid rods which orientation this angles describe. As the result the complete transformation is: $\hat{U}^T_{A,1}=\hat{U}^R_{A,1}\hat{U}_{A,1}$. Such rotation transformation operator commutes with $\hat{H}_c$ and due to it can't change the evolution of the transformed states $\cite{Aha,May2}$. \section {Quantum Clocks Models} To construct the relativistic covariant formalism of quantum RFs it's necessary first to define the time in such RFs. In nonrelativistic mechanics time $t$ is universal and is independent of observer, while in relativistic case each observer in principle has its own proper time $\tau$. We don't know yet the nature of time , but phenomenologically it can be associated with the clock hands motion or some other relative motion of the system parts $\cite {Hol}$. In Special Relativity the time in moving frame $F^1$ can be defined by external observer at rest measuring the state of $F^1$ comoving clocks. We'll consider the same procedure in relativistic QM i.e. some clock observable being measured at some time from the rest frame gives the estimate of proper time of moving quantum RF $F^1$. For some clocks models $F^1$ internal evolution which define $F^1$ clocks motion and consequently its proper time $\tau_1$ can be factorized from $F^1$ c.m. motion. Its quantum c.m. motion described by the relativistic Schrodinger equation for massive boson. This is Klein-Gordon square root (KGR) equation in which only positive root will be regarded for initial positive energy state $\cite{Schw}$. Solving Dirack constraints it was shown recently that this first order equation is completely equivalent to free Field secondary quantization $\cite {Git2}$. For our relativistic model we should regard more strictly the features of reference frames and clocks, taking into account the internal motion. Consider the evolution of some system $F^1$ where the internal interactions described by the Hamiltonian $\hat{H}_c$ are nonrelativistic , which as was discussed in chap.1 is a reasonable approximation for the measuring devices or clocks. We'll use the parameter $\alpha_I=\frac{\bar{H}_c}{m_1}$ ,where $m_1$ is $F^1$ constituents total rest mass. In $F^1$ c.m. $\alpha_I=m_1^{-1}\langle\varphi_c|\hat{H}_c|\varphi_c\rangle$ where $\varphi_c$ is $F^1$ internal state of ($\ref{A1}$). It describes the relative strength of the internal $F^1$ interactions and for the realistic clocks is of the order $10^{-10}$. In addition we'll assume that all RF constituents spins and orbital momentums are compensated so that its total orbital momentum is zero, like in $\alpha$-particle ground state. In this case the system $F^1$ c.m. motion can be reduced to the motion of the spinless boson with the mass $m_1$ and in the next order the mass operator $m_t=m_1+H_c$ will be used. We'll start the proper time study with the simple models of quantum RFs with clocks, yet we expect its main results to be true also for the more sophisticated models. To introduce our main idea let's regard the dynamics of the moving clocks in Special relativity $\cite {Dew}$. We'll suppose that the proper (clocks) time is defined by the coordinate $\theta$ describing some internal system motion independent of its c.m. motion. For the simplicity assume that Hamiltonian of clocks $H_c$ results in the trajectory $\theta(t)=\omega t+\theta_0$ of the clocks canonical observable $\theta$, which renormalized into the time observable $\tau=\frac{\theta}{\omega}$. This is the property which is expected from ideal clocks and the simplest example of such system is the motion of free particle relative to observer $\tau=\frac{x}{v}$ $\cite {Hol}$. For this and some other clocks models described below the Hamitonian of clocks with mass $m_1$ which c.m. moves with momentum $\vec{p}_1$ relative to ARF : $$ H_T=(m_t^2+\vec{p}_1^2)^\frac{1}{2} $$ where $m_t=m_1+H_c$. If $\theta ,\vec{p}_1$ commutes, solving Hamilton equations in ARF time $\tau_0$ one obtains $\theta(\tau_0)=B_1\omega \tau_0+\theta_0'$, where $B_1=\frac{m_t}{H_T}$ coincides with Lorentz boost value. So as expected , if $\theta$ is measured by the observer at rest he finds the proper time $\tau_1=B_1 \tau_0$ of moving frame. Yet we'll show that the quantum fluctuations of RF motion results in the principally new additional effects. One of the most simple and illustrative quantum clocks models is the quantum rotator proposed by Peres \cite{Per}. The rotator Hamiltonian $\hat{H}_c=-2\pi \omega i \frac{\partial}{\partial\theta}$ , where $\theta$ is the rotator's polar angle. Preparing the special initial state $\varphi_c(\theta)=|v^0_J\rangle$ at $t=0$, where $J$ is its maximum orbital momentum one obtains the close resemblance of the classical clocks hand motion. The clocks state $\varphi_c(\theta-2\pi\omega t)$ for large $J$ has the sharp peak at $\bar{\theta}=2\pi\omega t$ with the uncertainty $\Delta_{\theta}=\pm \frac{\pi}{N}$ and can be visualized as the constant hand motion on the clocks circle. Our main clocks model - $C_x$ exploits the nonrelativistic particle motion relative to observer with Hamitonian $H_c=\frac{\vec{p}^{\,2}}{2m}$ \cite {Hol}. Let's consider the particle 3-dimensional motion, but choose as its initial state at $t=0$ the Gaussian packet factorized in $x$ direction which momentum state vector is : \begin {equation} \phi_c(\vec{p})=A \phi(p_y,p_z) e^{\frac{\sigma_x^2}{2}(\bar{p}_x -p_x)^2} \quad \label {CD} \end {equation} for which $\bar{p}_x\ne 0$. $\sigma_x$ is the initial wave packet spatial spread. Then the simplest Hermitian observable which gives the time estimate is $\hat{\tau}=\frac{m x}{\bar{p_x}}$ - the particle's position on the arbitrary $x$ axe. It describes the nonshifted measurement with $\bar{\tau}=t$ and the finite dispersion $D_0(t)$ for $0<t<\infty \cite {Hol}$. In fact in $C_x$ model $\hat{\tau}$ is the clocks hand position operator or the pseudotime operator, and not a time operator in a strict sense $\cite{Hol,Per}$. So from all sides $C_x$ can be regarded as the realistic clocks model in which measuring $\hat{\tau}$ one obtains the correct $t$ estimate with some statistical error having quantum origin. $C_x$ wave function $\varphi_c(x,t)$ evolution can be factorized as the packet centre of gravity motion with the constant velocity $\frac{p_x}{m}$ and the packet smearing around it. For the given initial state there is unambiguous correspondence between the state vector $|\varphi_c(x,t)\rangle$ and time $t$, so the quantum clocks synchronization at $t=0$ means the preparation of the state $\varphi_c(x,0)$. From the corresponding Heisenberg equation one can find Heisenberg position operator for the Hamiltonian $H_c$ : \begin {equation} x(t)=(\frac{p_xt}{m}+x_0) \label {CAC} \end {equation} where $x_0=x(0)$ is Schrodinger position operator If $\bar{x}_0=0$ the corresponding clock time operator, which will be extensively used in relativistic theory can be decomposed as : $$ \hat{\tau}=t+\frac{p_x-\bar{p}_x+x_0m}{\bar{p}_x} $$ The first term gives the time expectation value and the rest gives the clocks dispersion $D(t)$. To simplify our discussion we'll consider also the clocks model $C_0$ with the linear approximation of the position operator $x(t)=\omega t+x_0$ where parameter $\omega=\frac{\bar{p}_x}{m}$ which is the analog of Peres clocks for unbounded motion. $C_0$ Hamiltonian $H_c^0=\omega p_x$ is unbounded from below for the continuous spectra, but for the interpretation of the relativistic clock effects it's unimportant. Any initial $C_0$ state ($\ref{CD}$) evolves as $\varphi^0_c (x-\omega t)$, so the initial form of wave function is conserved and only its centre of gravity moves. Now we'll consider the relativistic $C_x$ model in which RF $F^1$ and the particle $G^2$ system $S_2$ motion is relativistic. We'll suppose that ARF proper time $\tau_0$ is defined also by some quantum clocks ,which dispersion is so small that can be neglected and $\tau_0$ is the parameter. If $F^1$ internal interactions neglected $F^1$ c.m. motion described by the massive boson wave packet evolution and $S_2$ Hamiltonian $H_T$ in ARF is the sum of two KGR Hamiltonians for the positive energy states $\cite{Schw,Git2}$: \begin {equation} H_T=(m_1^2+\vec{p}_{1}^{\,2})^{\frac{1}{2}}+(m_2^2+ \vec{p}_{2}^{\,2})^{\frac{1}{2}}=(s+\vec{p}_s^{\,2})^{\frac{1}{2}} \label {C0} \end {equation} , where $\vec{p}_s=\vec{p}_1+\vec{p}_2$ and $s$ is invariant mass square. $\sqrt{s}$ can be regarded as the Hamiltonian of two objects $G^2, F^1$ relative motion in their c.m.s. equal to system $S_2$ mass operator : \begin {equation} m_t= \sqrt{s}=(m^2_1+\vec{q}^{\,2})^{\frac{1}{2}} +(m^2_2+\vec{q}^{\,2})^{\frac{1}{2}} \label {C2XYY} \end {equation} where $\vec{q}$ is $G^2$ relative invariant momentum \cite{Coe}. If $|\bar{q}|$ is small we can choose as $p_x$ - clock momentum $\vec{q}$ projection along any suitable direction for which $\bar{q}_x\ne 0$. In this case $F^1 , G^2$ relative motion can be regarded as nonrelativistic and $F^1$ mass operator approximated : $$ m_t\simeq m_s+\frac{q_x^2}{2\mu_{12}}+E_k(q_y,q_z) $$ ,where $\mu_{12}$ is $G^1,F^2$ reduced mass, $m_s=m_1+m_2$ is $S_2$ rest mass. In this case $E_k$ is small and can be omitted in the calculations. Like in nonrelativistic case $F^1$ proper time in this $C_x$ relativistic model can be estimated measuring in ARF the distance $x=x_2-x_1$ between $F^1$ and the particle $G^2$ which operator is equal to : $x=i\frac{\partial}{\partial q_x}$. For the obtained $m_t$ $S_2$ Hamiltonian $H_T$ can be formally rewritten : \begin {equation} {H}_T=[(m_s+H_c)^2+\vec{p}_s^{\,2}]^{\frac{1}{2}} \label {CA11} \end {equation} where $H_c=\frac{q_x^2}{2\mu_{12}}$. Moreover it is reasonable to assume that this square root Hamitonian can describe the evolution of any clocks model with nonrelativistic interactions $H_c$ i.e. for $\alpha_I\ll 1$ $ \cite{Schw}$. Here and below the algebraic operations with the operators (if they don't result into singularities) means Tailor raw decomposition. If $F^1, G^2$ relative motion is nonrelativistic we can assume for the beginning that $F^1$ and $S_2$ c.m.s. proper time practically coincide. For the classical motion $F^1$ Lorentz factor in $S_2$ c.m.s. $(1+\frac{\vec{q}^{\,2}}{m^2})^{\frac{1}{2}}$ and below we'll show that in quantum case their difference is also negligible. It's impossible to resolve in analytical form the Schrodinger equation for $H_T$ of ($\ref {CA11}$) , only some approximated solutions discussed below can be found. $S_2$ observables evolution can be found solving Heisenberg equation for the Hamitonian $H_T$ of (\ref{CA11}) or for exact Hamitonian of (\ref{C0}) as will be done below \cite {Hol}. After the simple algebra one obtains $x$ evolution in ARF proper time $\tau_0$ : \begin {eqnarray} \dot{x}=-i[x,H_T]= \frac{ -im_t}{(m_t^2+\vec{p}_s^2)^\frac{1}{2}}[x,H_c]=-iB_1[x,H_c] \label {C0H} \end {eqnarray} We'll call the operator $B_1(\vec{p}_s,m_t)$ the time boost operator, which interpretation will be discussed after some calculations. The clock observables we obtain in this clock models are the functions of canonical momentums only and due to it their factor ordering is unimportant for our problem. After the commutators calculations we can approximate operator $m_t$ by the parameter $m_t\simeq m_s+\frac{\bar{q}_x^2}{2m}$. The operator $x$ easily restored from $\dot{x}$ : $$ x(\tau_0)=B_1(\vec{p}_s,m_t)\frac{q_x\tau_0}{\mu_{12}}+x_0 $$ where $x_0$ is Schroedinger position operator for $\tau_0=0$. If we take that $\bar{x}_0=0$ it results into $F^1$ proper time operator : \begin {equation} \hat{\tau}_1= B_1(\vec{p}_s,m_t) \frac{q_x}{\bar{q}_x}\tau_0 +\frac{\mu_{12} x_0}{\bar{q}_x} \label {C2XX} \end {equation} Its meaning will be discussed after some calculations, but formally it's $F^1$ moving clocks hand position measured in ARF at the moment $\tau_0$. $\tau_1$ operator in $C_0$ model have the simpler form which prompts its interpretation : \begin {equation} \hat{\tau}_1= B_1(\vec{p}_s,m_t) \tau_0 +\frac{ x'_0}{\omega} \label {CXT} \end {equation} If $\bar{x}'_0=0$ $C_0$ $\hat{\tau}_1$ expectation value $\bar{\tau}_1=\bar{B}_1\tau_0$ coincides with the classical Lorentz time boost value. Its dispersion have the form : \begin {eqnarray} D_{\tau}=D_L(\tau_0)+D_c=D_B\tau_0^2+\bar{D}_2\tau_0+D_0 \label {C2X} \end {eqnarray} where $D_B=\bar{B}^2_1-(\bar{B}_1)^2$ and $D_0=\langle\frac{x_0^{'2}}{\omega^2}\rangle$ is the clocks mechanism dispersion, which for $C_0$ is time independent. Operator $D_2$ is equal to : \begin {equation} D_2=\frac{B_1 x_0+x_0 B_1}{\omega} \label {C2ZZ} \end {equation} The numerical calculations show that for $C_0$ localized states $D_2$ expectation value is very small and can be neglected. If $D_0$ is small $\tau_1$ fluctuations are defined mainly by $D_L(\tau_0)$ Lorentz boost dispersion stipulated by $\vec{p}_s$ fluctuations in $F^1$ wave packet. It's independent of the clocks mechanism and demonstrates that the proper time measurement have the principal quantum uncertainty growing unrestrictedly proportional to $\tau_0^2$. For $C_x$ model the factor $\frac{q_x}{\bar{q_x}}$ in ($\ref{C2XX}$) produces additional $\hat{\tau}_1$ fluctuations. Due to it Lorentz boost expectation value differs only for the small factor of the order $\alpha_I$ : $$ \bar{\tau_1}=\tau_0\bar{B}_1[1+\frac{\bar{B}_1}{\sigma_x^2\mu_{12}m_s} (1-\bar{B}_1^2)] $$ It results from $m_t$ dependence on $p_x$ and reflects influence of clocks energy on total mass. We'll neglect this effect in $C_x$ dispersion also described by ansatz (\ref {C2X}), but with different parameters : \begin {eqnarray} D_2=\frac{\mu_{12}}{\bar{q}_x^2}(q_xB_1x_0+x_0q_xB_1) ; \\ D_B=\frac{\bar{q}_x^2}{(\bar{q}_x)^2}\bar{B}_1^2-(\bar{B}_1)^2 \label {C2Y} ; \quad D_0=\frac{ \mu_{12}^2 \sigma_x^2}{\bar{q}_x^2} \label {C2WW} \end {eqnarray} Here $\bar{D}_2=0$ for the gaussian wave packets (\ref{CD}) and any other localizable states. Due to $q_x$ fluctuations absent in $C_0$ model the part of $D(\tau_1)$ : $$ D_x= D_0 +\frac{\bar{q}_x^2-(\bar{q}_x)^2}{(\bar{q}_x)^2}(\bar{B}_1)^2\tau_0^2 $$ can be related to the packet smearing along $x$ coordinate, regarded as the clocks mechanism uncertainty. To illustrate the physical meaning of this time operator let's consider the corresponding approximate solutions of $F^1$ state evolution equation for Hamiltonian (\ref{CA11}). For $\alpha_I\rightarrow 0$ we can decompose $H_T$ of (\ref{CA11}) in the first $\alpha_I$ order : \begin {equation} -i\frac{d\Psi_s}{d\tau_0} \simeq [(m_s^2+\vec{p}^2_s)^{\frac{1}{2}}+ \frac{m_1\hat{H}_c}{(m^{2}_s+\vec{p}_s^2)^{\frac{1}{2}}}]\Psi_s \label {C01} \end {equation} Here the first term is independent of $H_c$ which permit to represent $\Psi_s$ as the sum of factorized states. The second term is in fact the product of clock Hamitonian and Lorentz boost $B_1$. Let's choose the initial $F^1$ state $\Psi_s(0)=\Phi_s(\vec{p}_s)\varphi_c(x,0)$ and $\Phi_s=\sum c_l|\vec{p}_{sl}\rangle$, where the sum denotes the integral over $\vec{p}_s$. From our definition of quantum clocks synchronization it follows that $\Psi_s(0)$ describes $F^1$ clocks synchronized with ARF clocks at $\tau_0=0$. Solving equation ($\ref{C01}$) one finds : \begin {equation} \Psi_s(\tau_0)=\sum c_l\varphi_c(x ,B_l\tau_0) |\vec{p}_{sl}\rangle e^{-iE(\vec{p}_{sl})\tau_0} \label {C0A} \end {equation} where $E(\vec{p})=(m_s^{2}+\vec{p}^2)^\frac{1}{2}$ , $ B_l=B_1(\vec{p}_{sl},m_s)$. For linear clock $C_0$ Hamiltonian $H_c=H_c^0$ for small $\alpha_I$ this state can be rewritten : \begin {equation} \Psi_s(\tau_0)=\sum_l c_l \varphi^0_c(x-\omega B_l\tau_0) ~|\vec{p}_{sl}\rangle e^{-iE(\vec{p}_{sl})\tau_0} \label {C0B} \end {equation} To make the situation more clear suppose that $\varphi^0_c(0)=\delta(x)$, which evolves at rest into $\delta(x-\omega \tau_0)$ . Then $x$ measurement defines the time $\tau$ of quantum clocks at rest unambiguously and with zero dispersion, but $\Psi_s$ of ($\ref {C0B}$) in general isn't $x$ eigenstate. It means that at any $\tau_0>0$ $\Psi_s$ is the entangled superposition of the states $\varphi_c^0$ which $F^1$ clocks acquires at the consequent $\tau_1$ moments. As was shown there is one-to one correspondence between clock state $\varphi_c(x,t)$ and the time moment $t$ and in some sense it can be regarded as the 'superposition' of $F^1$ proper time moments, or more precisely $F^1$ states existed at this moments. For example $F^1$ clocks hand can show 3,4 and 5 o'clocks simultaneously which can be tested by $x$ measurement at some $\tau_0$ in ARF. This spread corresponds to $D_B$ dispersion term resulting from the $F^1$ momentum $\vec{p}_s$ uncertainty. For the realistic clocks their $x$ dispersion given by $D_0$ isn't zero even at rest and this two terms added as statistically independent effects. $\Psi_s$ for $C_x$ Hamitonian is given by ($\ref{C0A}$) and admits the same interpretation. It corresponds to the more complicated form of time dependent dispersion ($\ref{C2WW}$) which can be eventually factorized into the same two parts - relativistic and clock mechanism. So we conclude that the interpretation which follows from the approximate Schrodinger equation agrees well with Heisenberg operator calculus. In fact operator $\tau_1$ describes $F^1$ proper time in the limit when this clock dispersion is very small and the clock energy is much less then $F^1$ total mass energy i.e. $\alpha_I\rightarrow 0$. Obtained results suppose that the proper time of any quantum RF being the parameter in it simultaneously will be the operator from the 'point of view' of other RF. Qualitatively the appearance of RF proper time fluctuations can be understood considering the superposition of momentum eigenstates $|\vec{p}_{si}\rangle$ in $S_2$ wave packet as the superposition of $S_2$ velocities $\vec{\beta}_i$ and corresponding Lorentz factors $\gamma_1(\vec{\beta}_i)$. In Special Relativity $F^1$ proper time $\tau_1$ measured at the same $\tau_0$ in ARF depends on $\gamma_1$. If we formally extends this dependence on $F^1$ wave packet motion we get that the proper time will fluctuate proportionally to $ \gamma_1$ spread. So $F^1$ clocks measurement in ARF shows how much time passed in $F^1$ in this particular event and can give the different value for another event of the same ensemble. It means that the time moments in different RFs corresponds only statistically with the dispersion $D_{\tau}$ in ARF given by (\ref {C2X}). It differs from Special Relativity where one to one correspondence between $\tau_1, \tau_0$ time moments always exists , but can be incorporated into relativistic QEP if we find the analogous time relations between two quantum RFs of finite mass. In fact $\tau_1$ is more correct to relate to $S_2$ c.m.s. rest frame, but regarding the difference between $F^1$ and $S_2$ c.m.s. proper time operators $\tau'_1,\tau_1$ it's easy to show that they coincide if $\bar{q}_x\rightarrow 0$. From it we conclude that the principal part of the relativistic time operator, independent of any particular clocks mechanism features have the form in the limit $\alpha_I\rightarrow 0$ : \begin {equation} \hat{\tau}'_1=B_1(\vec{p}_1,m_1)\tau_0 \label {CBBB} \end {equation} Moreover this formulae permits to define formally the time operator for any object including the single massive particle. This operator form of $\tau'_1$ is closely connected with Fock-Shwinger proper time $\tau_F$ formalism interpretation and will be discussed in detail in the forcoming paper $\cite{Fock,Schw}$. Note only that $\hat{\tau}'_1(\tau_0)$ measurement gives $F^1$ proper time $\tau_F$ estimate at $\tau_0$ moment of ARF time. On the opposite in Fock-Shwinger formalism $\tau_F$ is the parameter time to which particular values operators $\hat{\tau}_0(\tau_F), \vec{r}_1(\tau_F)$ related. In distinction with our formalism it makes $\tau_F$ interpretation confusing, because $\vec{r}_1$ and other $ F^1$ operators are measured in ARF, hence the time of measurement defined in $F^1$ to which as we have shown in quantum case they related only statistically. The practical realization of $x$ measurement in ARF can be the intricated procedure, which scheme we don't intend to discuss here. Note only that to perform it one should measure simultaneously the distance between $F^1$ and $G^2$ and their total momentum giving total velocity and this two operators commute. Some examples of the analogous nonlocal observables measurements are described in \cite{Aha2}. The most disputable question here is the relativistic particle coordinate measurements. Yet in the considered case, when the relative $F^1,G^2$ average velocity is small then $x$ is the nonrelativistic coordinate operator. Yet to prove the quantum equivalence principle it's necessary to perform the full relativistic calculations. We'll present such completely relativistic results for $C_x$ model using Newton-Wigner Hermitian operator of the space coordinate $\cite{Wig}$ which is the direct analog of nonrelativistic operator $x_1$ : \begin {equation} \hat{x}^1_{NW}=i\frac{d}{dp_{x1}} -i\frac{p_{x1}}{2(m_1^2+\vec{p}^2_1)} \label {C5} \end {equation} The operator of two objects relative coordinates conjugated to c.m. momentum $q_x$ can be derived from this objects c.m. Hamiltonian (\ref{C2XYY}) : \begin {equation} \hat{x}_{NW}=x+F(\vec{q})=i\frac{d}{dq_{x}} -i\frac{q_{x}}{\sqrt{s}}(\frac{1}{w_1}+\frac{1}{w_2}) \end {equation} where $w_i=(m_i^2+\vec{q}^2)^{\frac{1}{2}}$. The clocks time observable in $F^1$ rest frame is proportional to $x_{NW}$ : $$ \tau=\frac{x_{NW} - \bar{x}_{NW}(0)}{\bar{\beta}_x} $$ where $\beta_x=q_x(w_1^{-1}+w_2^{-1})$ is $F^2,G^1$ relative velocity, If we choose $\bar{x}_{NW}(0)=0$ , then solving Heisenberg equation in ARF for the Hamiltonian of (\ref{C0}) we find the resulting $F^1$ time operator : \begin {equation} \hat{\tau}_1= \frac{B_1(\vec{p}_s,m_t){\tau_0}\beta_x+x_{NW}(0)} {\bar{\beta}_x} \label {C5AA} \end {equation} where in $B_1$ $m_t=\sqrt{s}$. This is the exact relativistic expression for $\tau_1$ without assumption of $q_x$ smallness. $\bar{\tau}_1$ corresponds to Lorentz boost value $\bar{B}_1$ which depends both on $\langle \vec{p}_s \rangle$ and $\langle \vec{q} \rangle$. It's easy to note that the momentum dependent part of $x_{NW}$ is constant in time and consequently can only enlarge the clocks mechanism dispersion $D_0$. Due to it the dispersion structure is the same as for nonrelativistic relative motion of (\ref{C2WW}) but its members are described by the more complicated formulaes omitted here. In fact this calculations evidence that $x_{NW}$ meausurements introduces only additional time-independent clock dispersion of the order of $G^2$ Compton wavelength without changing our previous conclusions about time operator properties. In fact $F^1$ proper time measurement in ARF can be performed by two different methods which equivalence must be proved. In the first method described above the detector $D_0$ installed in ARF measures $\tau_1$ and induces $C_x$ state collapse. In the second one the detector $D_1$ installed in $F^1$ measures the clock state and after it $D_1$ signal transfered to ARF. In this case we should consider the collapse in the moving frame , which is difficult to describe. But we must note that independently of its mechanism such interaction happens after this clocks evolves to this state and so can't influence directly on their evolution, so it seems correct to neglect it at this stage. Obtained time-fluctuation effect reminds the well-known life-time dilatation for the relativistic unstable particles $\cite{Byc}$. In this framework such particle can be regarded as the elementary binary clock having only two states. Obtained results evidence that the proper time in Quantum RF depend on the RF quantum state, but doesn't prove QEP directly. To do it we must consider two finite mass RFs on equal ground and to find the time transformation between them. \section {Relativistic Quantum Frames} To calculate the time operator between two RFs of finite mass it's necessary first to find the particle evolution equation in quantum RF rest frame. In general the system Poincare group irreducible representations contain the information which permit to describe its evolution completely, but due to appearance of time operators to find this representations for quantum RF is quite a problem. Therefore we choose another way; first we'll find the free particle evolution equation and corresponding proper time operator from the phenomenological arguments. After it we'll find Poincare transformations for quantum RFs which confirm this Hamiltonian ansatz. We'll study here only restricted Hilbert space sector where RF and particles states has positive energy $\cite {Git2}$. Again we'll study the system $S_2$ of RF $F^1$ and the particle $G^2$ which momentums $\vec{p}_i$, energies $E_i$ are defined in ARF. In JFC formalism described above we choose ARF FC $\vec{p}^{\,2}_0\approx 0$ and $S_2$ Hamitonian : \begin {eqnarray} H=(m_0^2+\vec{p}_0^2)^\frac{1}{2}+ (m_1^2+\vec{p}^2_{1})^\frac{1}{2}+H_{2} \label {C2} \end {eqnarray} Like in nonrelativistic case quantization means that all $\vec{p}_i$ are the operators and state vector ascribed also to ARF $|\vec{p}_0=0\rangle$. This formalism reproduces all relativistic QM results for $m_0\rightarrow\infty$. Going to $F^1$ rest frame we'll assume that the proper time parameter $\tau_1$ can be defined in it from $F^1$ clocks measurements extrapolation as was described in previous chapter. Choosing $F^1$ FC $\vec{p}^{\,2}_{11}\approx 0$ from the correspondence principle we'll suppose that the momentums expectation values in $F^1$ rest frame are given by Lorentz transformations with velocity $\vec{\beta}=\frac{\vec{p}_1}{E_1}$. : \begin {eqnarray} \vec{p}_{1i}=\vec{p}_i+ \frac{(\vec{n}_1\vec{p}_i)(E_1-m_1)\vec{n}_1-E_i\vec{p}_1}{m_1} +\vec{F}_i(\vec{p}_0) \label {C20} \end {eqnarray} where $\vec{n}_1=\vec{p}_1 |\vec{p}_1|^{-1}$. $\vec{F}_i$ are undefined at this stage operators for which $\langle \vec{F}_{i}\rangle=0$ and can be neglected in the following calculations. If $G^2$ have spin zero then the Hamiltonian $H$ transformed from ARF to $F^1$ is equal : \begin {eqnarray} H^1_R=H^1_0+H^1_1+H^1_2= \sum_{i=0}^2 (m^2_i+\vec{p}^{\,2}_{1i})^\frac{1}{2} \label {CC2} \end {eqnarray} For classical Special Relativity where normally RF supposed to have infinite mass $\vec{p}_{1i}, H^1_R$ corresponds to the canonical momentums for finite mass RFs $\cite {Lan}$. We see that $F^0$ motion is factorized from $S_2$ Hamiltonian $H^1=H^1_1+H^1_2$, and so in $F^1$ proper time $\tau_1$ $S_2$ evolution equation is : \begin {equation} -i\frac{d\psi^1}{d\tau_1}= [(m_1^2+\vec{p}_{11}^{\,2})^\frac{1}{2}+ (m_2^2+\vec{p}^{\,2}_{12})^\frac{1}{2}]\psi^1 =(s(\vec{q})+\vec{p}_s^{\,2})^\frac{1}{2}\psi^1 \label {C4} \end {equation} where $S_2$ c.m. observables $\vec{q},\vec{p}_s$ defined in chap.3. Solutions of this equation describe $G^2$ normalized free wave packet localizable relative to $F^1$ rest frame : \begin {equation} \Psi^1(\tau_1)=\varphi_2(\vec{p}_{12}) e^{-iE^1\tau_1} |m_2 ,\vec{p}_{12}\rangle |m_1,\vec{p}_{11}=0\rangle=\\ \varphi'_2(\vec{q}) e^{-iE^1\tau_1} |\sqrt{s} ,\vec{p}_s=\vec{p}_{12}\rangle |m_1,-\vec{q}\rangle|m_2,\vec{q}\rangle \label {CD1} \end {equation} expressed also via $S_2$ c.m. observables. Here $E^1=E^1_1+E^1_2$ are $H^1$ eigenvalues. They differ from the standard KGR energy only on $m_1$ and so we can use in $F^1$ rest frame the standard KGR momentum spectral decomposition and the states scalar product $\cite{Schw}$. In $F^1$ rest frame together with its proper time $\tau_1$ the space coordinate can be defined. We choose arbitrarily as $G^2$ coordinate (nonhermitian) operator in $F^1$ : $\hat{x}_{12}=i\frac{\partial}{\partial q_{x}}$ and corresponding Hermitian Newton-Wigner operator can be easily derived. Note that $x_q$ defined in $F^1$ differs from the same operator defined in c.m.s., yet our following results doesn't depend on the particular form of this operator. $x_{12}$ also differs from the operator $x_{p}=i\frac{\partial}{\partial p_{12x}}$ which corresponds to the classical distance between $F^1$ and $G^2$. They coincide only in the limit $m_1\rightarrow\infty$ or in nonrelativistic case. Now we can calculate $F^2$ proper time operator as function of the proper time in $F^1$. To perform it we assume again that $F^2$ c.m. motion is equivalent to the spinless particle $G^2$ motion. In the described framework the Hamiltonian of $F^2$ with $C_0$ or $C_x$ clocks in $F^1$ rest frame can be obtained substituting in $\hat{H}^1$ of (\ref{C4}) $m_2=m'_{2}+\hat{H}_c$. $\hat{\tau}_2$ can be found solving Heisenberg equation for $F^2$ clocks coordinate $\dot{x}=-i[x,H^1]$ analogously to ($\ref{C0H}$). If we omit analogously to (\ref{CBBB}) the members describing the clocks mechanism fluctuations the $F^2$ proper time operator $\hat{\tau}_2$ is equal : \begin {equation} \hat{\tau}_2=\frac{m_2 \tau_1}{(m^2_2+\vec{p}_{12}^2)^\frac{1}{2}} \approx \frac{m'_2 \tau_1}{(m'^2_2+\vec{p}_{12}^2)^\frac{1}{2}} =\hat{B}_1 (\vec{p}_{12},m'_2)\tau_1 \label {C6} \end {equation} This formalism is completely symmetrical and the operator obtained from (\ref{C6}) exchanging indexes 1 and 2 relates the time $\hat{\tau_1}$ in $F^1$ and $F^2$ proper time - parameter $\tau_2$. The Special Relativity limit when $\tau_2$ becomes the parameter is obvious and analoguosly to it the average time boost depends on whether $F^1$ measures $F^2$ clocks observables, as we consider or vice versa, and this measurement makes $F^1$ and $F^2$ nonequivalent $\cite{Lan}$. The new effect will be found only when $F^1$ and $F^2$ will compare their initially synchronized clocks. In QM formalism this synchronization means that $F^2$ state prepared at the moment $\tau_0$ can be factorized as $\Phi_2(\vec{p}_{12}) \varphi_c(x,0)$ analogous to ($\ref{C0A}$). If this $F^2$ time measurements repeated several times (to perform quantum ensemble) it'll reveal not only classical Lorentz time boost , but also the statistical spread having quantum origin with the dispersion given in (\ref{C2X}). Obtained relation between two finite mass RFs proper times evidence that Quantum Equivalence principle can be correct also in relativistic case. If the number of particles $N_g>1$ then for the system state description the clasterization formalism can be used $\cite{Coe}$. According to it for $N=3$ Hamiltonian in $F^1$ of two free particles $G^2,G^3$ rewritten through the system canonical observables acquires the form : \begin {equation} \hat{H^1}= (m_1^2+\vec{p}_{11}^2)^\frac{1}{2}+(s_{23}+\vec{p}^2_{1,23})^\frac{1}{2} =(s+\vec{p}_s^2)^\frac{1}{2} \label {C7} \end {equation} ,where $\sqrt{s}_{23}$ is $G^2, G^3$ invariant mass, $\sqrt{s},\vec{p}_s$ are the system total invariant mass and momentum. In clasterization formalism at the first level the relative motion of $G^2, G^3$ defined by $\vec{q}_{23}$ their relative momentum is considered. At the second level we regard them as the single quasiparticle - cluster $C_{23}$ with mass $\sqrt{s}_{23}$ and momentum $\vec{q}$ in the system c.m.s. It transformed to $\vec{p}_{1,23}$ momentum in $F^1$ and so at any level we can regard the relative motion of two objects only. This procedure can be extended in the obvious inductive way to arbitrary $N$. If we have two reference frames $F^1,F^2$ and $N_g\ne0$ then their relative momentums can be also described by the cluster formalism. Due to appearance of the time operator between two RFs to find the states transformations operator which we denote $\hat{U}^s_{2,1}(\tau_2,\tau_1)$ is quite a problem and here we'll obtain it only phenomenologically for some simple examples. Consider first the case $N=2$ when $S_2$ include $F^1,F^2$ only and its state in $F^1$ rest frame $\Psi^1(\tau_1)$ is the solution (\ref{CD1}) of eq. (\ref{C4}). We'll take that it transformed by $U^F_{2,1}$ into state $\Psi^2(\tau_2)$ in $F^2$ rest frame. If $F^1,F^2$ clocks are synchronized at $\tau_1=\tau_2=0$ then for this time moment $\Psi^2(0)=\hat{U}^F_{2,1}(0,0)\Psi^1(0)$ and from $F^1, F^2$ symmetry it follows : $|\Psi^2(0)\rangle=\varphi'_1(\vec{p}_{21}) |m_2,\vec{p}_{22}=0\rangle|m_1,\vec{p}_{21}\rangle$. $F^{1,2}$ internal wave functions $\varphi^{1,2}_c(x,0)$ at $\tau_1=0$ are obviously invariant and so omitted here. Like in nonrelativistic case we introduce $\vec{p}_f=\vec{p}_{11}+\vec{p}_{12}$, $\vec{p}'_f=\vec{p}_{21}+\vec{p}_{22}$ and conjugated $\vec{r}_f, \vec{r}'_f$. From the correspondence with Lorentz transformations it should give $\langle\vec{p}_{12}\rangle=-\frac{m_2}{m_1}\langle\vec{p}_{21}\rangle$ and if to demand $\vec{q}_2=-\vec{q}$ then the simplest transformation is equal to : \begin {equation} \hat{ U}^F_{2,1}(0,0)=P_f e^{i a_f\vec{p}_f\vec{r}_f} \label{CBBW} \end {equation} where $a_f=\ln\frac{{m}_1}{m_2}$, $P_f$ is $\vec{r}_f$ reflection (parity) operator. We see $\hat{U}^F_{21}(0,0)$ ansatz practically coincides with nonrelativistic transform of (\ref{BBB}) for $N=1$. The passive $S_N$ transformation for spinless $G^i$ also found from the correspondance principle as the minimal extension of standard Poincare transformations : \begin {equation} \hat{U}^s_{2,1}(0,0)=U^F_{21}(0,0)\prod_{j=3}^N e^{-i\vec{\beta}_{f}\vec{N}'_j} \label {C08} \end {equation} where velocity operator $\vec{\beta}_{f}=\vec{p}_{f}(H^{1}_2)^{-1}$, $\vec{N}'_i=H^1_{i}\frac{\partial}{\partial\vec{p}_{1i}} +\frac{\partial}{\partial\vec{p}_{1i}}H^1_i$ are $G^i$ Poincare generators in $F^1$ which coincide with standard ansatz. Then the transformation operator for arbitrary $\tau_1,\tau_2$ is : \begin {equation} \hat{U}^s_{21}(\tau_1,\tau_2)= \hat{W}_2(\tau_2)\hat{U}^s_{21}(0,0)\hat{W}_1^{-1}(\tau_1) \label{C8} \end {equation} , where $\hat{W}_{1,2}(\tau_{1,2})=exp(-i\tau_{1,2}\hat{H}^{1,2})$ are $S_N$ evolution operators and $H^{1,2}$ - $S_N$ Hamiltonians in $F^1,F^2$ rest frames. It means that despite $\tau_2$ and $\tau_1$ are correlated only statistically through $\hat{\tau}_2$ nevertheless $S_N$ state vectors in $F^2, F^1$ at this moments are related unambiguously. Transformed $S_N$ momentums are : \begin {eqnarray} \vec{p}_{21}=-\frac{m_2}{m_1}\vec{p}_{12}+d_1\vec{p}_{11} \quad , \vec{p}_{22}=-\vec{p}_{11}+d_2\vec{p}_{11}\\ \vec{p}_{2i}=\vec{p}_{1i}+ \frac{(\vec{n}_{12}\vec{p}_{1i})(E^1_2-m_2)\vec{n}_{12}-E^1_i\vec{p}_{12}} {m_2} +d_i\vec{p}_{11} \label {C99} \end {eqnarray} , where $\vec{n}_{12}=\frac{\vec{p}_{12}}{|p_{12}|}$, $E^1_i$ are $G^i$ energies in $F^1$. If to demand that all relative momentums $\vec{q}_{ij}$ conserved (or reflected), then $d_i$ can be calculated, but due to their unimportance we omit it here. It's easy to see that $\vec{p}_{1i}$ of (\ref {C20}) for ARF to $F^1$ transform follows from $U^s_{2,1}$ after the simple substitutions, and so the semiqualitative Hamiltonian derivation of (\ref{C20}) was consistent. We see that the passive spinless $G^3$ transformation differs from the standard one only by the change of velocity parameter to the operator $\vec{\beta}$ which commutes with $G^3$ Hamiltonian. To present full Poincare group we must include $F^2$ rotations which were considered in $\cite {Aha,May}$. In brief for spinless $G^3$ the rotations generators $\vec{M}'=\vec{J}$ are the standard orbital momentums, but the parameters $\vec{\omega}_2$ or $\theta_2,\varphi_2$ are changed to operators as was explained in chap.2. For illustration we'll consider 2-dimensional case when in $F^1$ rest frame $F^1,F^2$ axes orientation are given by the operators $\theta_1\approx 0, \theta_2$ describing $F^1,F^2$ internal degrees of freedom. Consequently after acting by $U_{2,1}^s$ of ($\ref{C08}$) on the initial state one acts by the rotation operator $U_2^R$ so that complete transformation is : $$ U^T_{2,1}(0,0)=U^R_2 U^s_{2,1}(0,0)= U^{in}_2\prod_{l=1}^3 e^{i\theta_2 J'_{lz}} U^s_{2,1}(0,0) $$ where $J'_{lz}$ are $F^{1,2}$ or $G^3$ orbital momentums defined in $F^2$ rest frame. $U^{in}_2$ is the operator transforming internal coordinates $\theta_{1,2}$ given in $\cite {May2}$. Note that $U^R_2$ commutes with $G^3$ Hamitonian and so don't influence its dynamics. Due to the complications of $U^T_{21}$ general form calculations we can't yet to describe full passive Poincare group with $F^2$ rotations for $G^3$ arbitrary spin, so the obtained relativistic theory is in fact consistent only in 1-dimensional case. For 3-dimensions we can introduce the active spinless $G^3$ Poincare transformations in $F^1$ rest frame with parameters $\vec{a},vec{\beta},\vec{\omega}$ which have the same generators as the standard ansatz of $\cite {Schw}$. In general from $U^T_{21}$ ansatz we can expect that arbitrary spin $G^3$ irreducible representations and Poincare generators in $F^1$ rest frame coincides with the standard ones $\cite {Schw}$, yet this hypothesis must be proved. It was argued that RF quantum properties can become important in Quantum Gravity , where in principle one should quantize the field, matter and RF simultaneously $\cite{Rov,Unr,Bro}$. In principle our approach permits to calculate the time operator $\hat{\tau}_1$ for RF $F^1$ moving in the external gravitational field $g_{\mu\nu}(x)$. We assume that ARF is located in the region where this field is weak and so we can take $\tau_0=x_0$ - world time. Analogously to (\ref{CA11}) $F^1$ clocks Hamiltonian in ARF (for $g_{oa}=0$ gauge) : \begin {equation} \hat{H}_T=[g_{00}(m_1+H_c)^2+g_{00}g_{ab}p^a_1p^b_1] ^\frac{1}{2} \label{CG} \end {equation} ,where $a,b=1,3$ $\cite{Lan}$. Now $H_T$ depends on $x_{\mu}$ and due to it solving Heisenberg equation (\ref{C0H}) for the clocks hand coordinate $x_c$ one obtains the differential relation for $\tau_1$ : \begin {equation} d\hat{\tau}_1=\frac{\sqrt{g_{00}}(m_1+H_c)d\tau_0} {[(m_1+H_c)^2+g_{ab}p^a_1p^b_1] ^\frac{1}{2}}=\sqrt{g_{00}}B_g(x,\vec{p}_1)d\tau_0 \label {CGG} \end{equation} In this case $\hat{\tau}_1$ becomes the integral operator , where integral is taken over $\tau_0$ interval. If $g_{\mu\nu}$ is the classical metrics then this relation contains no new physics , except the additional gravitational 'red shift' time boost proportional to $\sqrt{g_{00}}$ $ \cite{Lan}$. But in Quantum Gravity $g_{\mu\nu}(x)$ becomes the operator and its fluctuations can induce the additional quantum fluctuations of the measured $F^1$ clocks time. Despite that this fluctuation calculations are quite complicated we can expect from the general Quantum Statistics rules $\cite{Hol}$ that they can be factorized from the considered Lorentz boost fluctuations induced by the $F^1$ momentum fluctuations : $$ D_T=D_G(\tau_0)+D_L(\tau_0)+D_o(\tau_0) $$ From this rules we can expect also that for $F^1$ motion in the homogeneous gravitation field $D_G$ will grows proportionally to $\tau_0$ analogous to QED fluctuations (Brownian motion effects). Note that this fluctuations must be independent of RF mass. This approach can give some new insight into the famous time problem of Quantum Gravity $\cite{Rov,Bro}$ which we discuss here briefly. In this aspect the situation in Classical and Quantum Gravity seems to differ principally. Strictly speaking if the metrics becomes the operator it stops to be space-time metrics which unambiguously defines the space-time geometry. Due to it the observer can correctly use only the operational definition of physical space-time by means of clocks and other measurements. In gravity this operational time can originate from some evolving observable of gravitation field or to be the operator describing the time measurement for some free matter object carrying some nongravitational 'foreign' clocks. The idea that space-time events can be described by their relation with some distributed system or media was extensively explored for long time $\cite {Dew}$ . The most close to our purposes is the incoherent dust system, each piece of it carrying clocks. Gravity ADM quantization for such system with selfgravitation account permits to extract positive Schrodinger hamiltonian as was shown by Brown and Kuchar $\cite{Bro}$. Hence the dust pieces motion in their model was described only semiclassically. Introduction of 'dust space' $\vec{z}$ permit to quantize the gravitation field. Yet the free quantum motion of dust pieces transforms $\vec{z}$ into the operator on the initial space-time manifold $x_{\mu}$ which makes this quantization procedure contradictory. We describe here briefly the model of dust RFs quantum motion where in the first approximatio its selfgravitation neglected. Let's consider first classical RF $F^1$ free falling in external gravitational field. In $F^1$ comoving 'Gaussian' frame where frame conditions imposed before the field variation we have $g'_{00}=1, g'_{0a}=0$. In this RF for the classical field gravity constraints fulfilled $H_a(x)=0, H_0(x)=0$ which permit to calculate $g'_{ab},p'_{ab}$ evolution for $F^1$ clock time solving corresponding Hamilton equations for $H_0$ $\cite {Bro}$. In quantum case this vacuum field constraints results in Wheeler - deWitt equation $\hat{H}_0\Psi=0$ from which Schrodinger Hamiltonian can't be derived easily. Now let's account RF quantum motion and suppose that this constraints holds true also in quantum $F^1$ comoving frame. $F^1$ proper time for the external observer is given by the operator analogous to ($\ref{CGG}$), but in comoving frame $\tau_1$ is just the parameter. In this case we can calculate field observables evolution in $F^1$ clocks time from Heisenberg equations for $H_0$ vacuum constraint : $$ \dot{g}'_{ab}(x)=-i[g'_{ab}(x),H_0(x)] $$ where the commutator in general is nonzero. Note that this equation is obviously local, so to calculate $g'_{ab}(x,\tau_1)$ we must define $g'_{ab},p'_{ab}$ only on a small spacelike surface region around $x$ at a preceding moment $\tau_1-d\tau_1$. Space coordinates $x_a$ supposedly can be defined at least in the close vicinity of $F^1$ analoguosly to the definition given above for the flat space-time. Obviously this approach have many associated problems some of which are the construction of multifingered time for quantum RF dust and the field theoretical behavior of such commutators, despite it seems to deserve additional study. For the conclusion we can claim that the extrapolation of QM laws on free macroscopic objects regarded as RFs prompt to change the common approach to the space-time which was taken copiously from Classical Physics. In this paper the relativistic covariant theory of quantum RFs constructed and at least in flat space-time it agrees with the principle of equivalence for quantum RFs. The quantum RF momentum uncertainty results in the quantum statistical fluctuations of Lorentz boost which relates the proper times in two RFs. So in this model each observer has its proper time - parameter and euclidian coordinate space which can't be related unambiguously with the another observers space-time and in this sense is local. \end{sloppypar} {\footnotesize
\subsubsection{The Origin and Resiliency of Potential Models} A central puzzle in hadron spectroscopy is the apparent absence of low energy degrees of freedom beyond those which can be attributed to the valence quarks ({\it e.g.}, gluonic or sea quark excitations). Very closely related to this puzzle is the apparent unimportance of strong meson loop corrections. A simple resolution of this puzzle has been proposed~\cite{adiabatic}. In the flux tube model~\cite{IsgPat}, the quark potential model arises from an adiabatic approximation to the gluonic and extra $q \bar q$ degrees of freedom embodied in the flux tube. This physics has an analog at short distances where perturbation theory applies. There $N_f$ types of light $q \bar q$ pairs shift (in lowest order) the coefficient of the Coulombic potential from $\alpha_s^{(0)}(Q^2)=\frac{12\pi}{33 {\it ln}(Q^2/\Lambda_0^2)}$ to $\alpha_s^{(N_f)}(Q^2)= \frac{12\pi}{(33-2N_f) {\it ln}(Q^2/\Lambda_{N_f}^2)}$, the net effect of such pairs thus being to produce a {\it new} effective short distance $Q\bar Q$ potential. Similarly, when pairs bubble up in the flux tube ({\it i.e.}, when the flux tube breaks to create a $Q\bar q$ plus $q\bar Q$ system and then ``heals" back to $Q\bar Q$), their net effect is to cause a shift $\Delta E_{N_f}(r)$ in the ground state gluonic energy which in turn produces a new long-range effective $Q\bar Q$ potential. It has indeed been shown~\cite{GIonV} that the net long-distance effect of the bubbles is to create a new string tension $b_{_{N_f}}$ ({\it i.e.}, that the potential remains linear). Since this string tension is to be associated with the observed string tension, after renormalization {\it pair creation has no effect on the long-distance structure of the quark model in the adiabatic approximation}. Thus the net effect of mass shifts from pair creation is much smaller than one would na\"\i{ve}ly expect from the magnitude of typical hadronic widths: such shifts can only arise from nonadiabatic effects \cite{NIonNonadiabatic}. It should be emphasized that no simple truncation of the set of all meson loop graphs can reproduce such results: to recover the adiabatic approximation requires summing over large towers of $Q\bar q$ plus $q\bar Q$ intermediate states to saturate their duality with $q \bar q$ loop diagrams which have strength at high energy. \subsubsection{The Survival of the OZI Rule} There is another puzzle of hadronic dynamics which is reminiscent of this one: the success of the OZI rule~\cite{OZI}. A generic OZI-violating amplitude $A_{OZI}$ can be shown to vanish like $1/N_c$, and this is often quoted as a rationale for the OZI rule. However, there are several unsatisfactory features of this ``solution" to the OZI mixing problem~\cite{LipkinOZI}. Consider $\omega$-$\phi$ mixing as an example. This mixing receives a contribution from the virtual hadronic loop process $\omega \rightarrow K \bar K \rightarrow \phi$, both steps of which are OZI-allowed, and each of which scales with $N_c$ like $\Gamma^{1/2} \sim N_c^{-1/2}$. The large $N_c$ result that this OZI-violating amplitude behaves like $N_c^{-1}$ is thus not peculiar to large $N_c$: it just arises from ``unitarity" in the sense that the real and imaginary parts of a generic hadronic loop diagram will have the same dependence on $N_c$. The usual interpretation of the OZI rule in this case ~-~-~-~ that ``double hairpin graphs" are dramatically suppressed ~-~-~-~ is untenable in the light of these OZI-allowed loop diagrams. They expose the deficiency of the large $N_c$ argument since $A_{OZI} \sim \Gamma$ is {\it not} a good representation of the OZI rule. (Continuing to use $\omega$-$\phi$ mixing as an example, we note that $m_\omega - m_\phi$ is numerically comparable to a typical hadronic width, so the large $N_c$ result would predict an $\omega$-$\phi$ mixing angle of order unity in contrast to the observed pattern of very weak mixing which implies that $A_{OZI} << \Gamma <<m$.) Unquenching the quark model thus endangers the na\"\i{ve} quark model's agreement with the OZI rule. It has been shown~\cite{GIonOZI} how this disaster is naturally averted in the flux tube model through a ``miraculous" set of cancellations between mesonic loop diagrams consisting of apparently unrelated sets of mesons ({\it e.g.}, the $K\bar K$, $K\bar K^*+K^*\bar K$, and $K^*\bar K^*$ loops tend to strongly cancel against loops containing a $K$ or $K^*$ plus one of the four strange mesons of the $L=1$ meson nonets). Of course the ``miracle" occurs for a good reason: the sum of {\it all} hadronic loops is dual to a closed $q \bar q$ loop created and destroyed by a $^3P_0$ operator \cite{3P0mesons,KI}, but in the closure approximation such an operator cannot create mixing in other than a scalar channel. It can also be shown \cite{GIonsbars} that current matrix elements like $\bar s \gamma^{\mu} s$ vanish in this same approximation. \subsubsection{A Summary Comment on Modelling the Effects of $q \bar q$ Pairs} The preceding discussion strongly suggests that models which have not addressed the effects of unquenching on spectroscopy and the OZI rule should be viewed very skeptically as models of the effects of the $q \bar q$ sea on hadron structure: large towers of mesonic loops are required to understand how quarkonium spectroscopy and the OZI rule survive once strong pair creation is turned on. In particular, while pion and kaon loops (which tend to break the closure approximation due to their exceptional masses) have a special role to play, they will not allow a satisfactory solution to these fundamental problems associated with unquenching the quark model and so cannot be expected to provide a reliable guide to the physics of $q \bar q$ pairs. Indeed, I hope the reader can appreciate just on the basis of this lightning review that there are great dangers in drawing conclusions about the strength, structure, or significance of $q \bar q$ pairs in hadrons from any model that has not dealt with these issues. \section {Unquenching Heavy Quark Decay} \subsection {Background} To unquench predictions of the quark model for semileptonic heavy quark decay, I will apply without alteration the model of Refs. \cite{GIonV,GIonOZI,GIonsbars} which solves the phenomenological problems associated with unquenching the quark model. In particular, I assume that the $q \bar q$ pair is created by the action of a pair creation Hamiltonian density $H_{pc}(x)$ in the $Q \bar d$ flux tube. I further assume that the pair is created with a nonlocality (corresponding to a finite constituent quark radius) in the coordinate $\vec v$. See Fig. 2. \bigskip \begin{center} ~ \epsfxsize=3.5in \epsfbox{Qqqdcoords.ps} \vspace*{0.1in} ~ \end{center} \noindent{ Fig. 2: The coordinates for a) the $Q \bar d$ system, and b) the $Q \bar q q \bar d$ system. In each diagram the cross $\times$ denotes the location of the center-of-mass; most results presented in this paper are in the heavy quark limit where the center-of-mass coincides with the position of $Q$.} \bigskip The coordinates used here in $Q \bar d$ are the standard center-of-mass and relative coordinates \begin{eqnarray} \vec R&=&{{m_Q \vec r_Q+m_d \vec r_{\bar d}} \over m_{Q \bar d}} \simeq \vec r_Q \label{eq:defR} \\ \vec r&=&\vec r_{\bar d}-\vec r_Q \end{eqnarray} while in $Q \bar q q \bar d$ the choice is \begin{eqnarray} \vec R'&=&{{m_Q \vec r_Q+m_q(\vec r_q+\vec r_{\bar q})+m_d \vec r_{\bar d}} \over m_{Q \bar q q\bar d}} \simeq \vec r_Q \label{eq:defR'} \\ \vec r&=&\vec r_{\bar d}-\vec r_Q \\ \vec v&=&\vec r_{\bar q}-\vec r_q \\ \vec w&=&{1 \over 2}(\vec r_q+\vec r_{\bar q})-{{m_Q \vec r_Q+m_d \vec r_{\bar d}} \over m_{Q \bar d}} \simeq {1 \over 2}(\vec r_q+\vec r_{\bar q})-\vec r_Q~, \end{eqnarray} where $m_{ij...k} \equiv m_i+m_j+...+m_k$. (Note that the three-vector coordinate $\vec w$ should not be confused with the Lorentz invariant heavy quark scalar product $w=v' \cdot v$ called ``double $u$" (or, in many European countries, ``double $v$") used in heavy quark form factors). Inverting, we have in $Q \bar d$ \begin{eqnarray} \vec r_Q&=&\vec R_{cm}-\epsilon_{d/Q\bar d}~ \vec r \simeq \vec R_{cm} \\ \vec r_{\bar d}&=&\vec R_{cm}+\epsilon_{Q/Q\bar d} ~\vec r \simeq \vec R_{cm}+\vec r \end{eqnarray} where $\epsilon_{\alpha/\beta} \equiv m_{\alpha}/m_{\beta}$. In $Q \bar q q \bar d$ we have \begin{eqnarray} \vec r_Q&=&\vec R'_{cm}-\epsilon_{q\bar q/Q\bar q q \bar d} ~\vec w -\epsilon_{d/Q\bar d} ~\vec r \simeq \vec R'_{cm} \\ \vec r_{\bar q}&=&\vec R'_{cm}+\epsilon_{Q\bar d/Q\bar q q \bar d} ~\vec w + {1 \over 2} \vec v \simeq \vec R'_{cm}+ \vec w + {1 \over 2} \vec v \\ \vec r_q&=&\vec R'_{cm}+\epsilon_{Q\bar d/Q\bar q q \bar d} ~\vec w - {1 \over 2} \vec v \simeq \vec R'_{cm}+ \vec w - {1 \over 2} \vec v \\ \vec r_{\bar d}&=&\vec R'_{cm}-\epsilon_{q\bar q/Q\bar q q \bar d}~ \vec w +\epsilon_{Q/Q\bar d}~ \vec r \simeq \vec R'_{cm}+\vec r~. \end{eqnarray} Note that most of the results of this paper are presented in the heavy quark limit where the approximations shown in these formulas will often be used. Thus for a $0^-$ state \begin{equation} \Psi_{Q \bar d} = {1 \over {(2 \pi)^{3\over 2}}} e^{i\vec P_{cm}\cdot \vec R} \psi_{Q \bar d}(\vec r~) \chi^0_{s_Q s_{\bar d}} \label{eq:Qbard} \end{equation} where $\chi^0_{s_Q s_{\bar d}}$ is the spin zero wavefunction, so that \begin{equation} \Phi_{Q \bar d} = \delta^3(\vec P - \vec P_{cm}) \phi_{Q \bar d}(\vec p) \chi^0_{s_Q s_{\bar d}} \end{equation} is the momentum space $Q \bar d$ wavefunction with \begin{equation} \phi_{Q \bar d}(\vec p) \equiv {1 \over {(2 \pi)^{3\over 2}}} \int d^3r e^{-i\vec p\cdot \vec r} \psi_{Q \bar d}(\vec r~) \end{equation} and accordingly \begin{equation} \vert P_{Q \bar d}(\vec P_{cm}) \rangle= \sqrt{2m_{Q \bar d}} \int d^3p \phi_{Q \bar d}(\vec p~) \chi^0_{s_Q s_{\bar d}} \vert Q(\epsilon_{Q/Q \bar d}\vec P_{cm}-\vec p, s_Q) \bar d(\epsilon_{d/Q \bar d}\vec P_{cm}+\vec p, s_{\bar d}) \rangle~. \label{eq:defP2} \end{equation} Note that in the limit $\vec P_{cm} \rightarrow 0$ relevant here, the factor $\sqrt{2m_{Q \bar d}} \simeq \sqrt{2E_{P_{Q \bar d}}}$ is purely conventional as $m_Q \rightarrow \infty$. When the flux-tube-breaking pair creation Hamiltonian ${\bf H}_{pc}^{q \bar q} \equiv \int d^3x H_{pc}^{q \bar q}(0, \vec x)$ acts, \begin{equation} {\bf H}_{pc}^{q \bar q} \vert P_{Q \bar d}(\vec P_{cm}) \rangle= \eta_{q \bar q}\vert P_{Q \bar q q \bar d}(\vec P_{cm}) \rangle \label{eq:Hpc} \end{equation} where, according to the flux tube model, \begin{enumerate} \item since the $q \bar q$ pair is created in the flux tube, its center-of-mass is found in a wavefunction $\psi_{ft}(\vec w, \vec r~)$ defined by the flux tube spatial profile, \item the {\it internal} wavefunction of the $q \bar q$ pair has $J^{PC}=0^{++}$ and, independent of $\vec w$ and $\vec r$, is of the form $\psi^m_{pc}(\vec v) \cdot \chi^{-m}_{s_q s_{\bar q}}$ where $\psi^m_{pc}(\vec v)$ has $L=1$, $\chi^{-m}_{s_q s_{\bar q}}$ has $S=1$, and $\psi^m \cdot \chi^{-m} \equiv {1 \over {\sqrt 3}}(\psi^1 \chi^{-1}-\psi^0 \chi^0+\psi^{-1} \chi^1)$, and \item the amplitude to find the $Q \bar d$ subsystem inside the $Q \bar q q \bar d$ system at relative separation $\vec r$ is identical to that in the ground state, namely $\psi_{Q \bar d}(\vec r~)$, since the pair creation Hamiltonian density acts locally and instantaneously on the flux tube. \end{enumerate} \noindent In this formulation, $\eta_{q \bar q}$ defines the strength of the pair creation, and the normalized state $\vert P_{Q \bar q q \bar d}(\vec P_{cm}) \rangle$ is determined by the wavefunction just described: \begin{equation} \Psi_{Q \bar q q \bar d} = {1 \over {(2 \pi)^{3\over 2}}} e^{i\vec P_{cm}\cdot \vec R'} \psi_{ft}(\vec w, \vec r~) \psi^m_{pc}(\vec v) \cdot \chi^{-m}_{s_q s_{\bar q}} \psi_{Q \bar d}(\vec r~) \chi^0_{s_Q s_{\bar d}} \end{equation} The component parts of this wavefunction and of Eq. (\ref{eq:Qbard}) are defined by previous studies. From ISGW2 \cite{ISGW2}, \begin {equation} \psi_{Q \bar d} \simeq {\beta^{3/2}_{Q \bar d} \over \pi^{3/4}}e^{-{1 \over 2}\beta^2_{Q \bar d}r^2} \end {equation} where $\beta_{Q \bar d}=0.41$ GeV as $m_Q \rightarrow \infty$ as determined variationally from a Coulomb-plus-linear-plus-hyperfine Schrodinger equation. The pair creation wavefunction $\vec \psi_{pc}$ is constrained in Ref. \cite{GIonV} by fitting decay data assuming the form \begin {equation} \vec \psi_{pc} \sim \vec v e^{-3v^2/8r^2_q} \equiv \vec v e^{-{1 \over 2} \beta^2_{pc} v^2} \label{eq:psipc} \end {equation} to have a quark radius $0 < r_q < 0.4$ fm. Given this constraint, I will take $r_q=0.3$ fm as a ``canonical" value, corresponding to $\beta_{pc} \simeq 0.58$ GeV, but will consider deviations of $\pm 0.1$ fm from this value as plausible. (This central value and range are guided by the difficulty of inventing a mechanism which could lead to a constituent quark radius $r_q < 0.2$ fm.) Ideally \cite{KI}, $\psi_{ft}(\vec w, \vec r~)$ should have a probability profile which is a tube around the $Q \bar d$ axis with ``caps" at $Q$ and $\bar d$. This structure is probably very significant for the decays of highly excited states, but since all our decays will emerge from the $Q \bar d$ ground state, I adopt a simpler and more heuristic model which simply takes \begin {equation} \psi_{ft}(\vec w)={\beta^{3/2}_{ft} \over \pi^{3/4}}e^{-{1 \over 2}\beta^2_{ft}w^2} \label{eq:psift} \end {equation} with $\beta^2_{ft}=fb$, $b$ being the string tension and $f$ a coefficient with ``canonical" value $2$ and an uncertainty estimated to be $\pm 1$ based on the calculations of the properties of the ground state wavefunction of the flux tube presented in Appendix A of Ref. \cite{KI}. Some defects in this simplification and some subtleties associated with both the alternative of using a flux-tube shape and nonrelativistic kinematics are discussed in Appendix A of this paper. Since strong decay amplitudes are determined by matrix elements of $H^{q \bar q}_{pc}(t, \vec x)$ between the decaying particle and the continuum, $\eta_{q \bar q}$ can be determined empirically. In Appendix B, I extract from $D^* \rightarrow D \pi$ and $K^* \rightarrow K \pi$ decays the coefficients $\eta_{q \bar q}$ for $q=u$ or $d$. For concreteness, I will assume following Refs. \cite{KI,GIonV}, that $\eta_{s \bar s}$ is identical, but that the $\eta_{Q \bar Q}$ for $Q=c,b,t$ are all zero. Finally, for ease of exposition, I will treat explicitly the case of a single $q \bar q$ flavor in what follows, but take into account $q=u,d,s$ in numerical results. Within these approximations, \begin{equation} \Phi_{Q \bar q q \bar d} = \delta^3(\vec P- \vec P_{cm}) \phi_{ft}(\vec \omega) \phi^m_{pc}(\vec \pi) \cdot \chi^{-m}_{s_q s_{\bar q}} \phi_{Q \bar d}(\vec p~) \chi^0_{s_Q s_{\bar d}} \end{equation} where $\phi_{ft}$, $\phi^m_{pc}$, and $\phi_{Q \bar d}$ are the momentum space wavefunctions corresponding to $\psi_{ft}$, $\psi^m_{pc}$, and $\psi_{Q \bar d}$, respectively, so that \begin{eqnarray} \vert P_{Q \bar q q \bar d}(\vec P_{cm}) \rangle &=& \sqrt{2m_{Q \bar q q \bar d}} \int d^3\omega \int d^3\pi \int d^3p ~\phi_{ft}(\vec \omega) \phi^m_{pc}(\vec \pi) \cdot \chi^{-m}_{s_q s_{\bar q}} \phi_{Q \bar d}(\vec p~) \chi^0_{s_Q s_{\bar d}} \nonumber \\ &&\vert Q(\epsilon_{Q/Q \bar q q \bar d}\vec P_{cm}-\epsilon_{Q/Q \bar d}\vec \omega - \vec p, s_Q) \bar q(\epsilon_{q/Q \bar q q \bar d}\vec P_{cm}+{1 \over 2}\vec \omega + \pi, s_{\bar q}) \nonumber \\ && q(\epsilon_{q/Q \bar q q \bar d}\vec P_{cm}+{1 \over 2}\vec \omega - \pi, s_q) \bar d(\epsilon_{d/Q \bar q q \bar d}\vec P_{cm}-\epsilon_{d/Q \bar d}\vec \omega + \vec p, s_{\bar d}) \rangle~. \label{eq:defP4} \end{eqnarray} (Up to this point, I have retained the exact kinematics of the nonrelativistic limit for finite $m_Q$, but from now on I will generally simplify results by taking the heavy quark symmetry limit $m_Q \rightarrow \infty$ with $\vec V_{cm} \equiv \vec P_{cm}/m_Q$ fixed.) While Eq. (\ref{eq:Hpc}) defines the action of ${\bf H^{q \bar q}_{pc}}$ on the $Q \bar d$ sector, it does not of course provide us with the ${\bf H^{q \bar q}_{pc}}$-perturbed ground state. This state is of the form \begin{equation} \vert P_{Q}(\vec P_{cm}) \rangle= {1 \over {\sqrt{1+c^2_{q \bar q}}}} \Bigl[ \vert P_{Q \bar d}(\vec P_{cm}) \rangle +c_{q \bar q}\vert \tilde P_{Q \bar q q \bar d}(\vec P_{cm}) \rangle \Bigr] \label{eq:mixedstate} \end{equation} where $\vert \tilde P_{Q \bar q q \bar d}(\vec P_{cm}) \rangle$ is a normalized $Q \bar q q \bar d$ state of the same general form as Eq. (\ref{eq:defP4}) but with a wavefunction $\tilde \Phi_{Q \bar q q \bar d}$ to be specified below and determined via \begin{eqnarray} c_{q \bar q}\vert \tilde P_{Q \bar q q \bar d}(\vec P_{cm}) \rangle &=& \sum_{ab} \int {d^3q \over 2m_{Q \bar q q \bar d}} \nonumber \\ &&{ {\vert \vec P_{cm}; ab(\vec q~) \rangle \langle \vec P_{cm}; ab(\vec q~) \vert {\bf H^{q \bar q}_{pc}} \vert P_{Q \bar d}(\vec P_{cm}) \rangle} \over {E_{ab}(\vec P_{cm},\vec q~)-E_{P_{Q \bar d}}(\vec P_{cm})} }~~~. \label{eq:deftildeP} \end{eqnarray} Here $\vert \vec P_{cm}; ab(\vec q~) \rangle$ is the two meson eigenstate with $Q \bar q$ in internal state $a$, $q \bar d$ in internal state $b$, and with $(Q \bar q)_a$ and $(q \bar d)_b$ having relative momentum $\vec q$ and total momentum $\vec P_{cm}$; $E_{ab}(\vec P_{cm}, \vec q~)$ and $E_{P_{Q \bar d}}(\vec P_{cm})$ are the total energies of the respective $Q \bar q q \bar d$ and $Q \bar d$ states with fixed total center-of-mass momentum $\vec P_{cm}$. To obtain a rough expression for $\vert \tilde P_{Q \bar q q \bar d} \rangle$, I make use of the approximate duality between each of the towers of states $a$ and $b$ and their corresponding free particle spectra in the internal relative momenta $\vec p_{Q\bar q}$ and $\vec p_{q\bar d}$, respectively. {\it E.g.}, I use \begin{equation} \sum_{a} \langle (Q\bar q)_a \vert \simeq \sum_{s_Qs_{\bar q}} \int d^3p_{Q\bar q}\langle Q(-\vec p_{Q\bar q}, s_Q)\bar q(\vec p_{Q\bar q},s_{\bar q}) \vert \label{eq:freeduality} \end{equation} where for simplicity I have illustrated the duality equation in the $Q \bar q$ center of mass frame. While this replacement is imperfect for low $a$ (corresponding to small $\vec p_{Q\bar q}$), since ${\bf H^{q \bar q}_{pc}}$ is quite pointlike, Eq. (\ref{eq:deftildeP}) has most of its strength for relatively massive states $a$ and $b$ and the use of duality should be satisfactory for our purposes. With these approximations we can change variables from $a,b, \vec q$ to $\vec p_{Q\bar q}$, $\vec p_{q\bar d},\vec q$, and then convert both sides of Eq. (\ref{eq:deftildeP}) to the variables $\vec \omega$, $\vec \pi$, and $\vec p$ to identify \begin {equation} c_{q \bar q} \tilde \Phi_{Q \bar q q \bar d} \simeq {{\eta_{q \bar q} \Phi_{Q \bar q q \bar d}} \over {\Delta E}} \label{eq:defcqbarq} \end {equation} where in the rest frame $\vec P_{cm}=\vec 0$ and with $m_Q \rightarrow \infty$, \begin {equation} \Delta E = 2m_q+{\pi^2 \over m_q}+{\omega^2 \over 4m_q}+\delta \end {equation} with \begin {equation} \delta=m_Q+m_d+{p^2 \over 2m_d}-m_{P_{Q\bar d}}~. \end {equation} In the duality approximation we have adopted, and with the wavefunction of the $Q \bar d$ system identical in $Q \bar q q \bar d$ and $Q \bar d$, $\delta \simeq 0$. Moreover, using the wavefunctions (\ref{eq:psipc}) and (\ref{eq:psift}) and the parameters given earlier \begin {equation} { <{ \omega^2 \over 4m_q}> \over <{ \pi^2 \over m_q}>} \simeq {3 \beta^2_{ft} \over 20 \beta^2_{pc}} \end {equation} is small so that \begin {equation} \Delta E \simeq 2m_q+{ \pi^2 \over m_q} \end {equation} and we may deduce that \begin{equation} \tilde \Phi_{Q \bar q q \bar d} \simeq \delta^3(\vec P- \vec P_{cm}) \phi_{ft}(\vec \omega) \tilde \phi^m_{pc}(\vec \pi) \cdot \chi^{-m}_{s_q s_{\bar q}} \phi_{Q \bar d}(\vec p~) \chi^0_{s_Q s_{\bar d}} \end{equation} {\it i.e.}, that $\tilde \Phi_{Q\bar q q \bar d}$ differs from $\Phi_{Q\bar q q \bar d}$ simply by the replacement $\phi_{pc} \rightarrow \tilde \phi_{pc}$ where \begin{equation} \tilde \phi^m_{pc}(\vec \pi) \equiv n^{-1/2} { {\phi_{pc}^m(\vec \pi)} \over {1+\pi^2/2m_q^2} }~. \label{eq:phipc} \end{equation} The normalization factor $n$ is given by \begin {equation} n=\int d^3\pi \vert { {\phi_{pc}^m(\vec \pi)} \over {1+\pi^2/2m_q^2} } \vert^2~. \end {equation} which may be quite well approximated by the formula $n={{1+y}\over{1+21y+12y^3}}$ with $y={\beta^2_{pc} \over 8m_q^2}$. From Eqs. (\ref{eq:defcqbarq}) and (\ref{eq:phipc}) follows the key relation that \begin {equation} c_{q \bar q}={ {n^{1/2} \eta_{q \bar q}} \over 2m_q }~~~. \label{eq:candeta} \end {equation} Note that $\tilde \phi^m_{pc}(\vec \pi)$ is softer than $\phi^m_{pc}(\vec \pi)$. We will often use a harmonic approximation \begin{equation} \tilde \phi^m_{pc}(\vec \pi)={{\pi^m} \over \pi^{3/4} \tilde \beta_{pc}^{5/2}}e^{-\pi^2/2 \tilde \beta^2_{pc}} \label{eq:harmonictildephi} \end{equation} to $\tilde \phi^m_{pc}$. It is of course the softer shape of $\tilde \phi^m_{pc}$ that makes $n<1$, and thus $\tilde \beta_{pc}$ can be determined by requiring that Eqs. (\ref{eq:phipc}) and (\ref{eq:harmonictildephi}) match near $\vec \pi=0$, {\it i.e.}, that \begin{equation} \tilde \beta_{pc}=n^{1/5} \beta_{pc}~~. \end{equation} With realistic parameter values, $n^{1/5} \sim 0.7$, so the softening is not dramatic: with our canonical value for $\beta_{pc}$, $\tilde \beta_{pc} \simeq 0.4$ GeV. \subsection{The Unquenched Isgur-Wise Function} We are now in a position to calculate the unquenched quark model contribution to the Isgur-Wise function \cite{IWoriginal}. By heavy quark symmetry, the form factors for a general $Q_1 \rightarrow Q_2$ transition can be calculated as matrix elements for the simpler $Q \rightarrow Q$ transition with an arbitrary current $\bar Q \Gamma Q$. We therefore focus on the matrix elements of the scalar current $\bar Q Q$ between $Q$-containing states: \begin {eqnarray} \xi^{QM}(w)&=& {1 \over 2m_{Q }}\langle P_Q({\vec P_{cm} \over 2}) \vert \bar Q Q \vert P_Q(-{\vec P_{cm} \over 2}) \rangle \\ &=& {1 \over {1+c_{q \bar q}^2}} [ \xi_{Q \bar d}^{QM}(w)+c_{q \bar q}^2 \xi_{Q \bar q q \bar d}^{QM}(w) ] \label{eq:xiqm} \end {eqnarray} where $w \equiv v' \cdot v \simeq 1+P_{cm}^2/2m^2_{Q } = 1+V_{cm}^2/2$ and where, \begin {eqnarray} \xi^{QM}_{Q \bar d}(w)&=& {1 \over 2m_{Q }}\langle P_{Q \bar d}({\vec P_{cm} \over 2}) \vert \bar Q Q\vert P_{Q \bar d}(-{\vec P_{cm} \over 2}) \rangle \\ &=& \int d^3r \psi^*_{Q \bar d}(\vec r~)e^{-i m_d \vec V_{cm} \cdot \vec r}\psi_{Q \bar d}(\vec r~) \label{eq:xi2} \end {eqnarray} and \begin {eqnarray} \xi^{QM}_{Q \bar q q \bar d}(w)&=& {1 \over 2m_{Q}}\langle P_{Q \bar q q \bar d}({\vec P_{cm} \over 2}) \vert \bar Q Q \vert P_{Q \bar q q \bar d}(-{\vec P_{cm} \over 2}) \rangle \\ &=& \xi^{QM}_{Q \bar d}(w) \xi^{QM}_{ft}(w) \label{eq:xi4} \end {eqnarray} with \begin {equation} \xi^{QM}_{ft}(w) \equiv \int d^3w \psi^*_{ft}(\vec w)e^{-2i m_q \vec V_{cm} \cdot \vec w}\psi_{ft}(\vec w)~. \label{eq:xift} \end {equation} (In these equations the notation ``$QM$" reminds us that we are calculating in the quark model so that the perturbative matching of these HQET matrix elements to field theory must be done at the quark model scale $\mu_{QM} \sim 1$ GeV.) {\it We see from Eq. (\ref{eq:xift}) that $\xi^{QM}_{Q \bar q q \bar d}$ does not depend on the poorly known $q \bar q$ wavefunction $\psi_{pc}$}. This simplification arises because this $q \bar q$ wave function defines the relative position of the $q$ and $\bar q$, while $\xi^{QM}_{ft}$ is sensitive only to the $q \bar q$ system's wave function relative to $Q$. Defining for $w-1 <<1$ \begin {eqnarray} \xi^{QM} & \simeq & 1- \rho^2_{wf}(w-1) \\ \xi^{QM}_{Q \bar d} & \simeq & 1- \rho^2_{Q \bar d}(w-1) \\ \xi^{QM}_{ft} & \simeq & 1- \rho^2_{ \bar q q }(w-1) \end {eqnarray} and recalling the conventional definition \begin {equation} \xi \simeq 1- \rho^2( w-1) \end {equation} we therefore have (displaying now explicitly the effects of summing over $q=u$, $d$, and $s$) simply \begin {equation} \rho^2_{wf}= \rho^2_{Q \bar d} +{ {\sum_q c^2_{q \bar q} \rho^2_{ \bar q q }} \over {1+\sum_q c^2_{q \bar q}} } \end {equation} where $\rho^2_{wf}$ is the nonrelativistic wavefunction contribution to $\rho^2$, {\it i.e.}, it excludes the relativistic $1 \over 4$ and the contribution $\Delta \rho^2_{pert}$ from matching to the low energy effective theory \cite{ISGW2}. Using Eqs. (\ref{eq:xi2}) and (\ref{eq:xi4}) we then have the old result \begin {equation} \rho^2_{Q \bar d}={m_d^2<r^2> \over 3}={m_d^2 \over 2 \beta^2_{Q \bar d}} \label{eq:rhoQbard} \end {equation} and the new correction from $q \bar q$ pairs \begin {equation} \rho^2_{ \bar q q }= {4m_q^2<w^2> \over 3}= {2m_q^2 \over \beta^2_{ft}} \end {equation} so that in the $SU(3)$ limit where $m_u=m_d=m_s=m_q$ \begin {eqnarray} \rho^2_{wf}&=& {m_d^2 \over 2 \beta^2_{Q \bar d}} +{ {\sum_q c^2_{q \bar q} {2m_q^2 \over \beta^2_{ft}}} \over {1+\sum_q c^2_{q \bar q}} } = {m_d^2 \over 2 \beta^2_{Q \bar d}} +{ { c^2_{q \bar q} } \over {1+3 c^2_{q \bar q}} } \Bigl({6m_q^2 \over \beta^2_{ft}}\Bigr) \\ & \equiv & {m_d^2 \over 2 \beta^2_{Q \bar d}} + \Delta \rho^2_{sea} ~. \label{eq:mainresult} \end {eqnarray} This is one of our main new results. It shows that even if the $c_{q \bar q}^2$ are large, the contribution of pairs to $\rho^2$ will be small in the adiabatic limit where they are highly localized in the flux tube ({\it i.e.}, as $\beta^2_{ft} \rightarrow \infty$). See Appendix A for a discussion of this result for a more general flux-tube shape. We will see next that to the extent that pairs contribute to the exclusive ``elastic" slope $\rho^2$, they will also contribute to the {\it inclusive} nonresonant semileptonic rate. \subsection{A Duality Interpretation of $\Delta \rho^2_{sea}$ via Bjorken's Sum Rule} \subsubsection{Motivation} We have just seen that even if $P_{Q_1}$ is full of $q \bar q$ pairs, they may not contribute to $\rho^2$. We will now see that it is incorrect to associate the relative probabilities of $Q_1 \bar d$ and $Q_1 \bar q q \bar d$ in $P_{Q_1}$ with the resonant and nonresonant parts, respectively, of $Q_1 \rightarrow Q_2$ semileptonic decay. As heavy quark symmetry requires, at $w=1$ the $Q_1 \rightarrow Q_2$ transition creates only $P_{Q_2}$ and $V_{Q_2}$ of the ground state ${s'}_{\ell}^{{\pi '}_{\ell}}= {1 \over 2}^-$ multiplet {\it independent of the structure of the ``brown muck"}. {\it I.e.}, in this limit the $Q_2 \bar d$ and $Q_2 \bar q q \bar d$ components of the hadronic final state, no matter what their relative strengths, form perfectly into the resonant states $P_{Q_2}$ and $V_{Q_2}$. For $w-1$ small but nonzero, Bjorken's sum rule \cite{Bj,IWonBj} tells us that the loss of rate from the ``elastic" transitions $P_{Q_1} \rightarrow P_{Q_2}$ and $P_{Q_1} \rightarrow V_{Q_2}$ relative to structureless hadrons with $\rho^2=1/4$ will be exactly compensated by the production of ${s'}_{\ell}^{{\pi '}_{\ell}}= {1 \over 2}^+$ and ${3 \over 2}^+$ states. In the valence quark model, this rate must appear in $Q_2 \bar d$ excited states. In Ref. \cite{IWonBj} this valence quark model duality to the quark level semileptonic decay was explicitly demonstrated. Here I will show that the $Q_1 \bar q q \bar d$ content of $P_{Q_1}$ leads in general to the production of both resonant and nonresonant final states, with the latter rates proportional to $\Delta \rho^2_{sea}$. In particular, in the adiabatic limit there will be {\it no} nonresonant production. To the extent that $\Delta \rho^2_{sea}$ is nonzero, nonresonant $(Q_2 \bar q)_a(q \bar d)_b$ final states with ${s'}_{\ell}^{{\pi '}_{\ell}}= {1 \over 2}^+$ and ${3 \over 2}^+$ will be produced to compensate for the additional loss of rate from the elastic channels which it causes. It is natural to expect the compensation to occur in these channels. The softening of the elastic form factors which depletes the rate to $P_{Q_2}$ and $V_{Q_2}$ will have its analog in inelastic resonance excitation form factors, so these rates will also be diminished and cannot compensate for the additional loss of rate from $P_{Q_2}$ and $V_{Q_2}$. The population of inelastic channels is thus the only avenue available for satisfying Bjorken's sum rule. Of course this is also intuitively appealing: the loss of rate to the elastic channels occurs because after the recoil from $-\vec P_{cm}/2$ to $+\vec P_{cm}/2$, the $q \bar q$ parts of the ground state wave functions of the initial and final states fail to overlap, and in so doing they must ``by conservation of probability" find themselves in {\it their} excited states, namely as $(Q_2 \bar q)_a(q \bar d)_b$ continua. We will now make these heuristic observations precise. \subsubsection{Production of Nonresonant States at Low $w-1$} \bigskip The $n^{th}$ valence state perturbed by ${\bf H_{pc}^{q \bar q}}$ (the generalization of Eq. (\ref{eq:mixedstate})) may be written \begin{equation} \vert M^{(n)}_Q(\vec P_{cm}) \rangle = cos \theta \vert M^{(n)}_{Q \bar d}(\vec P_{cm}) \rangle + sin \theta \vert X^{(n00)}_{Q \bar q q \bar d}(\vec P_{cm}) \rangle \end{equation} where $\vert M^{(n)}_Q\rangle$, $\vert M^{(n)}_{Q \bar d}\rangle$, and $\vert X^{(n00)}_{Q \bar q q \bar d}\rangle$ are the generalizations of the states $\vert P_Q \rangle \equiv \vert M^{(0)}_Q\rangle$, $\vert P_{Q \bar d}\rangle \equiv \vert M^{(0)}_{Q \bar d} \rangle$, and $\vert \tilde P_{Q \bar q q \bar d}\rangle \equiv \vert X^{(000)}_{Q \bar q q \bar d}\rangle$ of Eqs. (\ref{eq:mixedstate}), (\ref{eq:defP2}), and (\ref{eq:deftildeP}), respectively, and where under our assumptions that the state of the flux tube is independent of $n$ (the adiabatic approximation) and that $H_{pc}^{q \bar q}$ does not affect the coordinate $\vec r$, $\theta$ is independent of $n$. (The rationale for the notation $(n00)$ will become apparent below.) From this expression one can immediately obtain the generalization of our result for the elastic transition that \begin{eqnarray} \xi^{QM}(w)^{n'n} &\equiv&{1 \over 2m_Q} \langle M^{(n')}_Q(+{ \vec P_{cm} \over 2}) \vert \bar Q Q \vert M^{(n)}_Q(-{ \vec P_{cm} \over 2}) \rangle \\ &=& \xi^{QM}_{Q \bar d}(w)^{n'n}[cos^2 \theta + sin^2 \theta \xi^{QM}_{ft}(w)] \end{eqnarray} where $\xi^{QM}_{Q \bar d}(w)^{n'n}$ is the valence quark model generalization of the Isgur-Wise function for $n \rightarrow n'$ transitions and $\xi^{QM}_{ft}(w)$ is exactly the same $q \bar q$ overlap form factor that appears in the elastic $n=0 \rightarrow n'=0$ transition. Thus the unquenched result for small $w-1$ is \begin{equation} \xi^{QM}(w)^{n'n} = \xi^{QM}_{Q \bar d}(w)^{n'n}[1-sin^2 \theta \rho^2_{ \bar q q }(w-1)] \end{equation} as for the Isgur-Wise function with, once again, $sin^2 \theta={{\sum_q c^2_{q \bar q}} \over {1+\sum_q c^2_{q \bar q}}}$. Since the production of each inelastic resonant channel occurs with strength proportional to $w-1$ to a positive integral power, the $q \bar q$ modification of $\xi^{QM}(w)^{n'n}$ for $n'>0$ has no effect on the saturation of Bjorken's sum rule for $\xi(w)$ to order $w-1$ since it produces effects which are at least of order $(w-1)^2$. This is in accord with the expectations outlined above that the additional depletion of elastic rate by $q \bar q$ pairs must be compensated by the explicit production of a $(Q \bar q)_a(q \bar d)_b$ continuum. To see this we must introduce the states in the continuum orthogonal to $\vert M^{(n)}_Q \rangle$. To this end we define a complete set of states $\vert X^{(n \alpha \beta)}_{Q \bar q q \bar d}(\vec P_{cm}) \rangle$ in the $Q \bar q q \bar d$ sector. Here $n$, $\alpha$, and $ \beta $ are excitation quantum numbers associated with the $\vec r$, $\vec w$, and $\vec v$ coordinates, respectively. These states are {\it not} the eigenstates of this sector in the absence of ${\bf H_{pc}^{q \bar q}}$: the eigenstates are the $\vert \vec P_{cm};ab(\vec q~) \rangle$ defined above. However, we can expand \begin{equation} \vert \vec P_{cm};ab(\vec q~) \rangle = \sum_{n \alpha \beta} \phi_{ab}^{(n \alpha \beta)}(\vec q~)^* \vert X^{(n \alpha \beta)}_{Q \bar q q \bar d}(\vec P_{cm}) \rangle~. \end{equation} It follows that to lowest order in $\theta$ (or equivalently the pair creation operator) we can form an orthogonal set of ${\bf H_{pc}^{q \bar q}}$-perturbed states \begin{eqnarray} \vert M^{(n)}_Q(\vec P_{cm}) \rangle & \simeq & \vert M^{(n)}_{Q\bar d}(\vec P_{cm}) \rangle +\theta \sum_{ab} \int d^3q \phi_{ab}^{(n00)}(\vec q~)\vert \vec P_{cm};ab(\vec q~) \rangle \\ \vert X_{ab}(\vec P_{cm}, \vec q~) \rangle & \simeq & \vert \vec P_{cm};ab(\vec q~) \rangle -\theta \sum_n \phi_{ab}^{(n00)}(\vec q~)^* \vert M^{(n)}_{Q\bar d}(\vec P_{cm}) \rangle \end{eqnarray} where $(\alpha, \beta)=(0,0)$ define the universal state of the $q \bar q$ pair created in a flux tube by the action of ${\bf H_{pc}^{q \bar q}}$. Let us now compute the transition amplitude to the continuum: \begin{eqnarray} \langle X_{ab}(+{\vec P_{cm} \over 2}, \vec q~) \vert \bar Q Q \vert M^{(0)}_Q(-{\vec P_{cm} \over 2}) \rangle & \simeq & \theta[ \langle +{\vec P_{cm} \over 2};ab(\vec q~) \vert \bar Q Q \vert \tilde P_{Q \bar q q \bar d} (-{\vec P_{cm} \over 2}) \rangle \nonumber \\ ~~~~~&-& \sum_n \phi_{ab}^{(n00)}(\vec q~) \langle M^{(n)}_{Q \bar d}(+{\vec P_{cm} \over 2}) \vert \bar Q Q \vert M^{(0)}_{Q \bar d}(-{\vec P_{cm} \over 2}) \rangle] \\ && \nonumber \\ & \simeq & \theta[\sum_{n \alpha \beta} \phi_{ab}^{(n\alpha \beta)}(\vec q~) \langle X^{(n\alpha \beta)}_{Q \bar q q \bar d}(+{\vec P_{cm} \over 2}) \vert \bar Q Q \vert X^{(000)}_{Q \bar q q \bar d} (-{\vec P_{cm} \over 2}) \rangle \nonumber \\ ~~~~~&-& \sum_n \phi_{ab}^{(n00)}(\vec q~) \langle M^{(n)}_{Q \bar d}(+{\vec P_{cm} \over 2}) \vert \bar Q Q \vert M^{(0)}_{Q \bar d}(-{\vec P_{cm} \over 2}) \rangle]~. \label{eq:trans1} \end{eqnarray} As $m_Q \rightarrow \infty$, $ \vec P_{cm} = m_Q \vec V_{cm} $ is much larger than any internal momentum so the matrix elements $\langle Q \vert \bar Q Q \vert Q \rangle$ appearing in the $Q \bar d \rightarrow Q \bar d$ and $Q \bar q q \bar d \rightarrow Q \bar q q \bar d$ transitions here are identical. Moreover, since the $\bar Q Q$ current does not affect the internal state of the $q \bar q$ pair, $\beta$ is required to be zero. Thus Eq. (\ref{eq:trans1}) becomes \begin{equation} \langle X_{ab}(+{\vec P_{cm} \over 2}, \vec q~) \vert \bar Q Q \vert M^{(0)}_Q(-{\vec P_{cm} \over 2}) \rangle \simeq \theta \sum_n \xi^{QM}_{Q \bar d}(w)^{n0} \Bigl[\sum_{\alpha} \xi^{QM}_{ft}(w)^{\alpha 0} \phi_{ab}^{(n \alpha 0)}(\vec q~) - \phi_{ab}^{(n 0 0)}(\vec q~)\Bigr] \label{eq:trans2} \end{equation} where $\xi^{QM}_{ft}(w)^{\alpha 0}$ is the generalization of $\xi^{QM}_{ft}(w)$ encountered above, namely \begin{equation} \xi^{QM}_{ft}(w)^{\alpha 0} \equiv \int d^3w~ \psi^{(\alpha)}_{ft}(\vec w)^*e^{-2i m_q \vec V_{cm} \cdot \vec w} \psi^{(0)}_{ft}(\vec w) \end{equation} where $\psi^{(\alpha)}_{ft}$ is the $\alpha^{th}$ basis state for the expansion of the $q \bar q$ center of mass coordinate $\vec w$. Expanding the exponential in powers of $\vec V_{cm}$ as is appropriate for small $w-1$, we obtain Eq. (\ref{eq:xift}) of Section II for $\alpha=0$ and \begin{equation} \xi^{QM}_{ft}(w)^{\alpha \neq 0, 0} \simeq -2i m_q \vec V_{cm} \cdot \int d^3w~ \psi^{(\alpha)}_{ft}(\vec w)^* \vec w \psi^{(0)}_{ft}(\vec w)~. \end{equation} Since $\xi^{QM}_{ft}(w)^{0 0} \simeq 1 - \rho^2_{ \bar q q }(w-1)$ with $w-1=V^2_{cm}/2$, to leading order in $V_{cm}$ the $\alpha=0$ term in Eq. (\ref{eq:trans2}) cancels with $\phi_{ab}^{(n00)}(\vec q~)$ to leave \begin{equation} \langle X_{ab}(+{\vec P_{cm} \over 2}, \vec q~) \vert \bar Q Q \vert M^{(0)}_Q(-{\vec P_{cm} \over 2}) \rangle \simeq \theta \sum_{n, \alpha \neq 0} \xi^{QM}_{Q \bar d}(w)^{n0} \xi^{QM}_{ft}(w)^{\alpha 0} \phi_{ab}^{(n \alpha 0)}(\vec q~)~. \label{eq:trans3} \end{equation} An immediate consequence of this relation is that in the adiabatic limit ($\beta_{ft} \rightarrow \infty$), $\xi^{QM}_{ft}(w)^{\alpha 0} \sim \beta_{ft}^{-1} \rightarrow 0$ for $\alpha \neq 0$ so we have explicitly demonstrated that there is no nonresonant production in this limit. Since we have for our discussion assumed that $\psi^{(0)}_{ft}(\vec w)$ has the form of a ground state harmonic oscillator wave function, it is natural to use a harmonic oscillator basis as the orthonormal expansion functions for the variable $\vec w$. Doing so, it follows that only three basis states give nonzero contributions to Eq. (\ref{eq:trans3}) to leading order in $\vec V_{cm}$ since $\vec w \psi^{(0)}_{ft}(\vec w)$ is proportional to the three $n_w=0$, $\ell_w=1$ harmonic oscillator wave functions: \begin{equation} \psi^{[n_w=0, \ell_w=1,i]}_{ft}(\vec w) = \sqrt{2}{\beta_{ft}^{5/2} \over \pi^{3/4}} w^i e^{-{1 \over 2}\beta_{ft}^2w^2} \end{equation} in a Cartesian basis, giving \begin{equation} \xi^{QM}_{ft}(\vec w)^{[n_w=0, \ell_w=1,i]0} = -{ {i \sqrt{2} m_q V^i_{cm}} \over {\beta_{ft}} } \end{equation} and thence \begin{equation} \langle X_{ab}(+{\vec P_{cm} \over 2}, \vec q~) \vert \bar Q Q \vert M^{(0)}_Q(-{\vec P_{cm} \over 2}) \rangle \simeq \theta \sum_{n, i} \xi^{QM}_{Q \bar d}(w)^{n0} \xi^{QM}_{ft}(\vec w)^{[n_w=0, \ell_w=1,i]0} \phi_{ab}^{(n [n_w=0, \ell_w=1,i] 0)}(\vec q~)~. \end{equation} Next we note that \begin{equation} \xi^{QM}_{Q \bar d}(w)^{n\neq0,0} \sim (w-1)^k \end{equation} where $k$ is a positive integer by Luke's Theorem \cite{Luke} so that to order $\vec V_{cm}$ \begin{eqnarray} \langle X_{ab}(+{\vec P_{cm} \over 2}, \vec q~) \vert \bar Q Q \vert M^{(0)}_Q(-{\vec P_{cm} \over 2}) \rangle & \simeq & \theta \xi^{QM}_{Q \bar d}(w) \sum_i \xi^{QM}_{ft}(\vec w)^{[n_w=0, \ell_w=1,i]0} \phi_{ab}^{(0 [n_w=0, \ell_w=1,i] 0)}(\vec q~) \\ & \simeq & -{ {i \theta \sqrt{2} m_q \vec V_{cm}} \over {\beta_{ft}} } \cdot \vec \phi_{ab}(\vec q~) \end{eqnarray} where I have introduced the notation $\vec \phi_{ab}(\vec q~)$ for the three-vector $\phi_{ab}^{(0 [n_w=0, \ell_w=1,i] 0)}(\vec q~)$. This key result has a simple interpretation. Nonresonant production at order $w-1$ requires that the current $\bar Q Q $ acting on $Q \bar q q \bar d$ create neither the ground state nor the $P$-wave resonances. However, it cannot excite the $q \bar q$ internal coordinate $\vec v$ and, if it excites $\vec r$ then to order $w-1$ it has just produced the $Q \bar q q \bar d$ component of either the ground state or the $P$-wave resonances. {\it Hence nonresonant production to order $w-1$ occurs purely by excitation of the $q \bar q$ coordinate $\vec w$ to $\ell_w=1$.} The factors $\vec \phi_{ab}(\vec q~)$ are simply the projections of these $\ell_w=1$ states onto the continuum eigenstates consisting of mesons $a$ and $b$ with relative momentum $\vec q$. We are now in a position to verify that the $\Delta \rho^2_{sea}$ contribution to the slope of the Isgur-Wise function is indeed compensated by the production of these continuum $(Q \bar q)_a (q \bar d)_b$ states. The probability for their production at $w$ is, up to $(w-1)^2$ corrections, \begin{equation} dP(P_Q \rightarrow {\rm{continuum}}) \simeq \sum_{ab} \int d^3q \vert \langle X_{ab}(+{\vec P_{cm} \over 2}, \vec q~) \vert \bar Q Q \vert M^{(0)}_Q(-{\vec P_{cm} \over 2}) \rangle \vert^2 \end{equation} which since $ \sum_{ab} \int d^3q \phi^i_{ab}(\vec q~) \phi^j_{ab}(\vec q~)=\delta^{ij}~, $ gives \begin{eqnarray} dP(P_Q \rightarrow {\rm{continuum}}) & \simeq & { {2 \theta^2 m_q^2 V^2_{cm} } \over \beta^2_{ft} } \\ & \simeq & {4m_q^2 \over \beta^2_{ft}} \theta^2(w-1) \label{eq:nrtotal} \end{eqnarray} for the flavor $q$. On the other hand, according to Eq. (\ref{eq:mainresult}), the contribution of flavor $q$ to the loss of elastic rate is \begin{equation} dP(P_Q \rightarrow P_Q+V_Q)_{Q \bar q q \bar d} \simeq -{4m_q^2 \over \beta^2_{ft}} \theta^2(w-1)~. \end{equation} The two rates match, explicitly demonstrating the connection of $\Delta \rho^2_{sea}$ to the nonresonant continuum. \subsubsection{Production of Exclusive Nonresonant States at Low $w-1$} \bigskip It remains to assess the fractional population of individual continuum channels inside of the total given by Eq. (\ref{eq:nrtotal}). To do this we must calculate $\vec \phi_{ab}(\vec q~)$, which from its definition is \begin{equation} \phi_{ab}^{(n \alpha \beta)}(\vec q~) \delta^3(\vec P~'_{cm}-\vec P_{cm})= \langle \vec P~'_{cm};ab(\vec q~) \vert X^{(n \alpha \beta)}_{Q \bar q q \bar d}(\vec P_{cm}) \rangle \end{equation} for the case $(n \alpha \beta)=(0 [n_w=0, \ell_w=1,i] 0)$. These calculations are straightforward, but would be quite tedious without the introduction of several tricks described in Appendix C. Results of the calculations of $\tau^{(m)}_{1/2}(w)$ and $\tau^{(p)}_{3/2}(w)$ for number of low-lying nonresonant channels are given in Table I. \section{Unquenching Heavy Quark Decay: Results} \bigskip We now turn to the quantitative evaluation of $\Delta \rho^2_{sea}$ (which is a reflection of total nonresonant production) and then to the distribution of these decays into exclusive nonresonant channels. \subsection{The Total Nonresonant Rate} \bigskip As shown in Appendix B, the light quark amplitudes $\eta_{u \bar u}=\eta_{d \bar d}$ can be determined from strong decays to be \begin{equation} \eta_{q \bar q} \simeq 0.9 ~{\rm GeV} \end{equation} with an uncertainty of a factor of two mainly arising from a strong dependence on the poorly known quantity $\beta_{pc}$. It follows from Eq. (\ref{eq:candeta}) that \begin{equation} c_{q \bar q} \simeq 0.5~. \end{equation} Assuming $SU(3)$ symmetry for the contribution to $\Delta \rho^2_{sea}$, one then obtains from Eq. (\ref{eq:mainresult}) \begin{equation} \Delta \rho^2_{sea} \simeq {1 \over 4}~, \end{equation} corresponding to an increase of $\rho^2$ from the value $0.74 \pm 0.05$ quoted in ISGW2 to a value near unity. Either value would be in reasonable agreement with measurements \cite{rho2}. {\it Via} the Bjorken sum rule, such a $\rho^2_{sea}$ would be consistent with the possibility discussed in the Introduction that $16 \pm 8 \%$ of the inclusive semileptonic $\bar B$ rate is in nonresonant channels. This reasonable quantitative correspondence between our calculated $\Delta \rho^2_{sea}$ and a possible experimental anomaly should not be taken too seriously. The missing non-$(D+D^*)$ rate attributed to nonresonant production might be due to an ISGW2 underestimate of excited resonance production \cite{Wolf}, or to an experimental overestimate of non-$(D+D^*)$ production. Moreover, while it is a ``canonical estimate", $\Delta \rho^2_{sea}$ is subject to very substantial uncertainties: see Table II. \subsection{The Rate to Low-Lying Exclusive Nonresonant Channels} \bigskip Using the amplitudes of Table I and the formulas of Appendix C, one can easily calculate the {\it fractions} of $\Delta \rho^2_{sea}$ due to the individual low-mass nonresonant channels shown in Table III. Note that the nonresonant rate is highly fragmented: none of the many channels tabulated account for more than $4\%$ of the inclusive nonresonant rate (and thus no more than about $1\%$ of the total semileptonic rate). These results are consistent with previous studies of single low energy pion emission using heavy quark chiral perturbation theory \cite{Qxipt}. Note also that the thirty final states considered here account for only about $40\%$ of the nonresonant rate, so that most of this rate resides in states $ab$ composed of highly excited resonances. There are several reasons why the results given in Table III must be interpreted with caution. The most prominent is simply that they depend roughly on $\tilde \beta_{pc}^{-5}$! Recalling that $\tilde \beta_{pc}=n^{1/5} \beta_{pc}$, our canonical value for $\tilde \beta_{pc}$ is $0.4$ GeV, but over the range of $\beta_{pc}$ allowed by Table II, it varies from its canonical value by $\pm 0.1$ GeV. Although the factor $n^{1/5}$ makes this range narrower than that of $\beta_{pc}$ itself, it still leaves us with an uncertainty of more than a factor of two on the basic unit for production of exclusive nonresonant channels. The second word of caution concerns large corrections to the $m_Q \rightarrow \infty$ limit studied here which arise for real $\bar B$ decay. The problem is phase space: in the heavy quark limit, all final states occur at the $w$ of the underlying quark production process, {\it i.e.}, the full mass range of the final state spectrum of states is negligible compared to the heavy quark energies even at low $w-1$. In actual $\bar B \rightarrow X_c \ell \bar \nu_{\ell}$ decays, each final state $X_c$ will have a Dalitz plot that is a shrinking fraction of the $b \rightarrow c \ell \bar \nu_{\ell}$ Dalitz plot as $m_{X_c} \rightarrow m_B$. Thus in the heavy quark limit the loss of rate from the elastic channels $D+D^*$ would be {\it locally compensated} in the variable $w$. For finite $m_b$ and $m_c$, however, a loss of elastic rate will still occur {\it via} $\Delta \rho^2_{sea}$, but the compensating channels will experience a delayed turn-on because of their thresholds; indeed, some processes which would have helped to compensate $\Delta \rho^2_{sea}$ will be kinematically forbidden. These phase space suppressions lead to $\Lambda_{QCD}/m_Q$-type corrections to the inclusive rate, and therefore also corrections to the accuracy of quark-hadron duality. However, it is known on general grounds from heavy quark symmetry and the operator product expansion (OPE) that, as the energy release $m_b-m_c \rightarrow \infty$, the leading corrections to the inclusive rate must be of order $\Lambda^2_{QCD}/m^2_Q$ \cite{inclorig,SVonInclusives}. The resolution of this apparent paradox has been discussed in Ref. \cite{NIonInclusives}: the OPE has a radius of convergence which does not include the region in which (significant) hadronic thresholds are turning on, so for real $b \rightarrow c$ transitions there can be $\Lambda_{QCD}/m_Q$ corrections, and they can be substantial. However, associated with these $\Lambda_{QCD}/m_Q$ threshold effects, which would diminish the integrated contribution of individual hadronic channels below the level required for perfect duality, are $\Lambda_{QCD}/m_Q$ corrections to the rate for the production of such channels which {\it enhance} their production once they are above threshold. These counterbalancing effects soften the breaking of duality: they are the precursors of the perfect cancellation of $\Lambda_{QCD}/m_Q$ effects that must occur as $m_b-m_c \rightarrow \infty$. How then should one make use of Table III for real $b \rightarrow c$ transitions? The amplitudes $\tau_{1/2}^{(m)}$ and $\tau_{3/2}^{(p)}$ shown are the leading order predictions for their respective channels. For the reasons just outlined, we can expect that they will be enhanced by $\Lambda_{QCD}/m_Q$ corrections which are expected to be of the order of $25-50\%$ \cite{NIonInclusives}. However, despite this effect, I believe that the dramatically reduced size of their Dalitz plots (relative to that for the underlying quark process $b \rightarrow c \ell \bar \nu_{\ell}$) will reduce the actual population of all nonresonant states well below that expected from $\Delta \rho^2_{sea}$. Indeed, each channel shown in Table III has a continuous spectrum of masses from its threshold up to masses exceeding $m_B$. Consider, for example, the $D+\rho$ channel. Its mass is given by $m_{D \rho}(q^2)={\sqrt {m_D^2+q^2}}+{\sqrt {m_{\rho}^2+q^2}}$, while its dominant $D$-wave production rate is proportional to $q^4 e^{-q^2/4 \tilde{ \bar \beta}^2}$, where $2 \tilde{ \bar \beta} \sim 1$ GeV. The production is therefore very weak at low masses where the available phase space is generous, and peaks at $m_{D \rho} \sim m_B$ where it vanishes. Given these basic kinematic facts, it is clear that even this simple exclusive channel will be produced at a rate far less than in the heavy quark limit, and that most of $\Delta \rho^2_{sea}$ will be uncompensated. The third column of Table III gives the phase space factors by which individual channels will be reduced in real $\bar B \rightarrow X_c \ell \bar \nu_{\ell}$ decays relative to the heavy quark limit; the fourth column gives a very rough estimate for the {\it net} suppression of each channel as a product of this phase space factor and a generous guess that, after $\Lambda_{QCD}/m_Q$ corrections, each channel has a compensatory increase of $50\%$. Considering that in general the untabulated channels of yet more highly excited states $ab$ will suffer even greater phase space suppressions, an overall diminution of the nonresonant rate by at least a factor of four seems likely. \section{Conclusions} \bigskip Although the successes of the valence quark model and the arguments of the large $N_c$ limit provide indications that sea quarks play a relatively minor role in hadronic physics, this hope is far from being justified by our current understanding. Some failures of the quark model ({\it e.g.}, the proton spin crisis) and the known existence of strong real and virtual decay channel couplings indeed make blithely ignoring the role of $q \bar q$ pairs both phenomenologically and theoretically untenable. In this work I have examined the influence of $q \bar q$ pairs on the simplest ``real" hadrons: heavy quark mesons like the $\bar B$. This study has led to a number of qualitative insights which I believe are quite general in nature. In earlier work on ``unquenching the quark model", the success of the valence quark model in spectroscopy was shown to have a possible basis in the validity of an adiabatic approximation. In this approximation, both the confining flux tube and the many $q \bar q$ pairs it generates remain in their adiabatically evolving ground state as the valence quarks move. In this work I have shown that the same approximation leads to valence quark and therefore resonance dominance of the simplest current matrix elements: $\bar Q_2 \Gamma Q_1$ matrix elements of heavy quark mesons. The physical picture behind our results is simple and appealing. According to heavy quark symmetry, at small recoil the $Q_1 \rightarrow Q_2$ decay of the $Q_1$ ground state $P_{Q_1}$ will lead with unit probability to $P_{Q_2}$ and $V_{Q_2}$ {\it no matter how complicated the QCD ``brown muck" might be}. This simple observation makes it clear that for $\bar Q_2 \Gamma Q_1$ matrix elements the issue is not the probability of $q \bar q$ pairs in $P_{Q_1}$ but rather how rapidly as $w-1$ increases these pairs fail to overlap with those in $P_{Q_2}$, $V_{Q_2}$, and the $Q_2 \bar d$ excited states. I have shown explicitly that in the adiabatic limit these overlaps are perfect so that only valence states (the resonances) are produced. Moreover, I showed that violations of the adiabatic approximation can be directly associated with the production of nonresonant states. Thus this study leads to a way of understanding how the valence quark model can be so successful {\it even though} hadrons are full of $q \bar q$ pairs. While they are quantitatively very crude, these calculations also have interesting consequences for real $\bar B$ semileptonic decays. First of all, they suggest that $\rho^2_{dyn} \equiv \rho^2-{1 \over 4}$ is composed of two comparable parts: $\rho^2_{resonant} \simeq {1 \over 2}$ and $\rho^2_{nonresonant} \simeq {1 \over 4}$, though we have stressed that this split of $\rho^2_{dyn}$ is very model-dependent. In the heavy quark limit, Bjorken's sum rule would then lead one to expect (using the central experimental value of the $(D+D^*)$ fraction) that roughly $24\%$ (with a $50\%$ error) of semileptonic decays go into resonances (both ordinary $Q \bar d$ mesons and $Q \bar d$ hybrids), and roughly $12\%$ (with an error of a factor of two) go into nonresonant states. Since the former states will have most of their strength in the $2.4-3.0$ GeV region, they will suffer, relative to the heavy quark limit, phase space suppression factors varying from only $0.75$ to $0.50$ over this range which may be fully compensated by the $\Lambda_{QCD}/m_Q$ enhancements required by asymptotic duality \cite{SVonInclusives,NIonInclusives}. In contrast, we have seen that nonresonant states are expected to populate very high masses peaking in strength near $m_B$ and so to suffer a substantial {\it net} suppression factor of at least ${1 \over 4}$. From this study I therefore expect that ${\buildrel < \over \sim} 5\%$ of $\bar B$ semileptonic decays will be nonresonant! As a corollary of this last observation, I note that if a $12\%$ nonresonant semileptonic fraction is required for duality but only a quarter of this is realized, then duality will fail from this effect alone by $\sim 10\%$, as anticipated in Ref. \cite{NIonInclusives}. There is, however, a minor inconsistency associated with this conclusion. Experiment requires that $36 \pm 6\%$ of $\bar B$ semileptonic decays be non-$(D+D^*)$ decays, in contrast to the $\sim 25\%$ we would have estimated from the preceeding. As mentioned earlier, this could simply mean that the ISGW2 model underpredicts the production of excited charm mesons \cite{Wolf} or that experiment has overestimated non-$(D+D^*)$ production. Determining whether this discrepancy is real will require a more quantitative calculation than this one (and probably additional experimental measurements as well). Detailed experimental studies of the structure of the hadronic final state in semileptonic $b \rightarrow c$ decays can therefore answer some fundamental questions about the role of $q \bar q$ pairs and about duality in strong QCD. A vital feature of these systems is that duality is underwritten by Bjorken's sum rule, requiring an exact and local duality between quark- and hadronic-level decays in the heavy quark limit. In particular, the experimental determination of the strength and structure of these nonresonant contributions would immediately test the conclusions reached here that these $q \bar q$ effects are highly suppressed in real $b \rightarrow c$ decays, that such decays extend to very high masses, and that they are highly fragmented into many small channels. Independent of the outcome, examining this problem in Nature's simplest hadronic system under the action of its simplest current (a heavy-to-heavy nonsinglet transition) should prove to be an excellent starting point for eventually understanding the $q \bar q$ sea in all strongly interacting matter. In particular, given the complexity of QCD, this seems an essential first step before tackling the problems of duality and nonresonant production in ordinary deep inelastic scattering. \vfill\eject {\noindent \bf APPENDIX A: Flux Tubes and A Critique of Nonrelativistic Kinematics} \bigskip It is not difficult to make the simplification $\psi_{ft}(\vec w, \vec r~)=\psi_{ft}(\vec w)$ of Eq. (\ref{eq:psift}) more flux-tube-like. In the case where $\psi_{ft}$ depends on $\vec w$ and $\vec r$, Eqs. (\ref{eq:xi4})-(\ref{eq:xift}) become \begin {eqnarray} \xi^{QM}_{Q \bar q q \bar d}(w)&=& {1 \over 2m_{Q}}\langle P_{Q \bar q q \bar d}({\vec P_{cm} \over 2}) \vert \bar Q Q \vert P_{Q \bar q q \bar d}(-{\vec P_{cm} \over 2}) \rangle \\ &=& \int d^3w d^3r \psi^*_{Q \bar d}(\vec r~) \psi^*_{ft}(\vec w, \vec r~) e^{-2i m_q \vec V_{cm} \cdot \vec w} e^{-i m_d \vec V_{cm} \cdot \vec r} \psi_{Q \bar d}(\vec r~) \psi_{ft}(\vec w, \vec r~) \label{eq:xi4wr} \end {eqnarray} or, defining \begin{equation} \xi^{QM}_{Q \bar q q \bar d}(w)=1-\rho^2_{Q \bar q q \bar d}(w-1) \end{equation} we have \begin{eqnarray} \rho^2_{Q \bar q q \bar d} &=& {1 \over 3} \int d^3w d^3r (2m_q \vec w + m_d \vec r~)^2\vert \psi_{Q \bar d}(\vec r~) \psi_{ft}(\vec w, \vec r~) \vert^2 \\ &=& {1 \over 3}\langle (2m_q \vec w + m_d \vec r~)^2 \rangle \label{eq:rhoqbarq} \end{eqnarray} which reduces to the simplified results of the text in the appropriate limits. Now consider a generic example of a more realistic $\psi_{ft}$ that has a flux tube's shape: \begin{equation} \psi_{ft}(\vec w, \vec r~)={\beta_{ft} \over \pi^{1/2} }e^{-{1 \over 2}\beta^2_{ft} w^2_{\perp}}t(\vec w \cdot \hat r) \end{equation} where $\hat r=\vec r/r$, $\vec w_{\perp}=\vec w - (\vec w \cdot \hat r)\hat r$, and $t(\vec w \cdot \hat r)$, which depends only on the longitudinal variable $\vec w \cdot \hat r$, is a normalized tube-like function. ({\it E.g.}, one might have $t={1 \over \sqrt{r}} \theta (\vec w \cdot \hat r) \theta (r-\vec w \cdot \hat r)$ to create a cylindrical wavefunction that is gaussian transverse to $\vec r$ and constant between $Q$ and $\bar d$.) With such a wavefunction \begin{equation} \rho^2_{Q \bar q q \bar d}={1 \over 3}[4m_q^2\langle w^2_{\perp}\rangle + \langle (2m_q \vec w \cdot \hat r + m_d r)^2 \rangle] \label{eq:pluslongw} \end{equation} The first term is as expected intuitively: it is the unchanged $\vec w_{\perp}$ part of the result of the text. One might also naively interpret the second term as the old $\vec r$ term plus a new longitudinal contribution due to the assumed spatial distribution of $\vec w \cdot \hat r$ in a tube-like configuration along $\vec r$. I believe that the physics is more subtle than this. Consider the origin of $\rho^2_{wf}$ in the nonrelativistic kinematics of our model. In the $Q \bar d$ sector (see Eq. (\ref{eq:defR})), $ \vec r_Q = - m_d \vec r/ m_Q$ in the center of mass frame as $m_Q \rightarrow \infty$, while in $Q \bar q q \bar d$ (see Eq. (\ref{eq:defR'})), $\vec r_Q \rightarrow - (2m_q \vec w + m_d \vec r~)/ m_Q $. Since nonrelativistically $\rho^2={1 \over 3}m_Q^2 \langle r_Q^2 \rangle$, we see that Eq. (\ref{eq:rhoQbard}) for $\rho^2_{Q \bar d}$ and Eq. (\ref{eq:rhoqbarq}) for $\rho^2_{Q \bar q q \bar d}$ are simply consequences of these nonrelativistic relations. To see the dangers of this approximation when the string tension and its renormalization are large compared to $m_q$, consider the calculation of the mass of a system of heavy quarks $Q$ and $\bar d$ at separation $\vec r$ connected by a renormalized flux tube, {\it i.e.}, of the state (\ref{eq:mixedstate}) of the text which has string tension $b_{N_f}$ since it has the appropriate admixture of $q \bar q$ pairs. If its mass were determined nonrelativistically one would obtain \begin{equation} M^{nr}-m_Q =m_d+{{2m_q c_{q \bar q}^2} \over {1+ c_{q \bar q}^2}} \end{equation} {\it i.e.}, the effective mass opposite $Q$ (against which it must recoil to conserve the position of the center of mass) would be the probability-weighted masses of the pure $ \bar d$ state and the $ \bar q q \bar d$ admixture. On adding interactions (both the diagonal potential $b_0r$ and the off-diagonal perturbation $H_{pc}^{q \bar q}$ which mixes $Q\bar d$ and $Q \bar q q \bar d$), we obtain the correct answer \begin{equation} M^{adiabatic}-m_Q =m_d+b_{N_f}r~~. \end{equation} Thus the mass $2m_q$ does not in this circumstance have an independent reality as indicated by the nonrelativistic kinematics just described: it is subsumed into the properly renormalized string tension. Indeed, it is an implicit assumption of the model for the $q \bar q$ pairs described in the text that the unquenched flux tube also behaves like a relativistic string: it should support (only) transverse waves moving with the speed of light. Thus while the $q \bar q$ pairs change the longitudinal distribution of $\vec r_Q $, {\it this effect is already described in the flux tube model by the mass $b_{N_f}r$ residing in the flux tube}: when $r$ increases, not only does $\bar d$ move, but so does the center of mass of the flux tube. It would therefore be double-counting to include the effect on $\langle r^2_Q \rangle$ of the $2m_q\vec w \cdot \hat r$ term of Eq. (\ref{eq:pluslongw}). I should hasten to add that the nonrelativistic quark model used in this paper does not {\it explicitly} take this effect into account. To do so would require a full treatment of the flux tube degrees of freedom ({\it versus} the adiabatic, potential-model approximation used here). Nevertheless, even the nonrelativistic quark model has undoubtedly already taken some of this effect into account implicitly by its choice of such free parameters as $m_d$. ({\it E.g.}, in many applications the constituent quark mass is effectively $m_d=m^0_d+{1 \over 2}b \langle r \rangle$). In contrast, there is no mechanism in the quark potential model to take into account transverse displacements of $Q$ relative to $\vec r$ ~~\cite{NIonTransverse}. These transverse displacements are the true degrees of freedom of the (quenched and unquenched) flux tubes and the reaction of $Q$ to them makes a non-potential-model-type transverse contribution to $\rho^2$. By renormalizing the string tension from $b_0 \rightarrow b_{N_f} < b_0$, $q \bar q$ pairs increase the longitudinal contribution to $\rho^2$ (at least in the nonrelativistic approximation ${1 \over 2}b r << m_d$). However, this increase is {\it not} compensated by nonresonant production, since it is this same string tension $b_{N_f} < b_0$ which controls the structure and thereby the production of excited resonances. {\it I.e.}, the longitudinal effect of the pairs is real, but it simply renormalizes resonance physics. The dynamics behind this balancing act, characteristic of the adiabatic approximation, can be seen by calculating the contribution of the flux tube to the Isgur-Wise function \cite{NIonTransverse}: since the flux tube has only transverse internal degrees of freedom, it has no impact on longitudinal overlap integrals over the total separation $\vec r$. Based on the preceding arguments, the transverse contributions to $\langle r^2_Q \rangle$ would also be those of a relativistic string with string tension $b_0$ or $b_{N_f}$ in the quenched and unquenched flux tubes, respectively. In contrast to the longitudinal contribution to $\langle r^2_Q \rangle$, the transverse contributions correspond to the reaction of $Q$ to real internal degrees of freedom, and these degrees of freedom (both gluonic and $q \bar q$) can be excited by the action of the $\bar Q \Gamma Q$ current. When acting on the pure $Q \bar d$ piece of the state, the current excites hybrid mesons which in the quenched limit exactly compensate for the loss of rate from the elastic channel due to transverse contributions to $\langle r^2_Q \rangle$~~\cite{NIonTransverse}. In the $Q \bar q q \bar d$ sector, the current could in principle excite either of the strings internal to mesons $(Q \bar q)_a$ or $(q \bar d)_b$, or it could excite the center of mass of the $q \bar q$ pair. In the quark potential model approximation to this latter process, which is of course the one of interest for this paper, one would recover the result shown in Eq. (\ref{eq:pluslongw}) less the longitudinal part of $\vec w$: \begin{eqnarray} \rho^2_{Q \bar q q \bar d}&=&{4m_q^2 \over 3}\langle w^2_{\perp}\rangle + {m_d^2 \over 3}\langle r^2 \rangle \\ &=&{4m_q^2 \over {3\beta^2_{ft}}} + {m_d^2 \over {3\beta^2_{Q\bar d}}}~. \label{eq:transverserho} \end{eqnarray} The first term of this formula differs from the expression of the text by the factor ${2 \over 3}$ corresponding to two of the three degrees of freedom of $\vec w$ being active. Note that for this picture to be consistent, the total transverse contribution to $\langle r^2_Q \rangle$ must be that of a relativistic string with string tension $b_{N_f}$; the decomposition into $Q \bar d$ and $Q \bar q q \bar d$ components is only useful in identifying the compensating channels required by Bjorken's sum rule. However, in the renormalized string picture $\langle r^2_Q \rangle$ of course depends just on $b_{N_f}$, while Eq. (\ref{eq:transverserho}) shows a contribution proportional to $2m_q$. It would be interesting to examine how the dynamics of pair creation in the flux tube leads to such a term. I speculate that the mechanism is the ``consumption" of a piece of flux tube of length $\Delta r \sim 2 m_q/b$ in a nonlocal pair creation process. In summary, in this Appendix I have described several subtleties in the description of $q \bar q$ pair creation in the flux tube model, and pointed out some interesting issues which arise in the physics of the renormalized flux tube. While I believe these matters are important conceptually and are worthy of further study, I am convinced that other uncertainties described in the text are of far greater impact numerically on our results. Given this and the great convenience of the spherical approximation, I therefore chose this simpler if less basic framework for the discussion of the text. \vfill\eject {\noindent \bf APPENDIX B: Determining $\eta_{q \bar q}$} \bigskip Eq. (\ref{eq:Hpc}) defines the action of the pair creation Hamiltonian on $\vert P_{Q \bar d} \rangle$. This perturbation not only produces pairs to make the eigenstate $\vert P_Q \rangle$ of Eq. (\ref{eq:mixedstate}), but also leads to strong decays. In particular, the projection of the state (\ref{eq:defP4}) onto the continuum states $\vert \vec P_{cm};ab(\vec q~) \rangle$ determines the $P_Q \rightarrow (Q \bar q)_a (q \bar d)_b$ coupling constants. By heavy quark symmetry \cite{IWspec}, the same dynamics determine the $P^*_Q \rightarrow (Q \bar q)_a (q \bar d)_b$ coupling constants, where $P^*_Q$ is the vector partner of the pseudoscalar state $P_Q$. In this Appendix I use these facts to determine the strength parameter $\eta_{q \bar q}$ of Eq. (\ref{eq:Hpc}) by comparing to the decays $P^*_Q \rightarrow P_Q \pi$. I begin with a practical matter. As $m_Q \rightarrow \infty$, the decays $P^*_Q \rightarrow P_Q \pi$ are forbidden since $P^*_Q$ and $P_Q$ become degenerate heavy quark spin partners. However, $\Gamma(K^* \rightarrow K \pi)$ is known, and $\Gamma(D^* \rightarrow D \pi)$ can be deduced if one assumes that the successful phenomenology of magnetic dipole decays can be extended to $\Gamma(D^* \rightarrow D \gamma)$. (The total width of the $D^*$ is so small that only decay branching fractions and not decay widths are known. If one takes the $K^* \rightarrow K \pi$ and $K^* \rightarrow K \gamma$ decays and scales them appropriately in $m_Q$ assuming that heavy quark scaling works all the way down to $m_s$, the observed $D^* \rightarrow D \pi$ and $D^* \rightarrow D \gamma$ branching ratios are explained nearly perfectly. This is another example of the often-noted fact that in many circumstances a strange quark behaves like heavy quark.) Since the branching ratio for $D^{*0} \rightarrow D^0 \pi^0$ is well-determined experimentally, I will use the value $\Gamma(D^{*0} \rightarrow D^0 \pi^0) \simeq 30$ keV deduced in this manner as input ``data". The calculations themselves are simple. If $g_0 D^{0\dagger}\partial_{\mu} \pi^0 D^{0* \mu}$ is the effective Lagrangian density for the decay leading to $\Gamma(D^{*0} \rightarrow D^0 \pi^0)= {{g^2q^3} \over {48 \pi m^2_{D^*}}}$, then in the center of mass, with the pion emitted with momentum $\vec q$ from a $D^*$ with polarization $+1$ along $\hat z$, \begin{eqnarray} -{{ig_0 q_+} \over {\sqrt{2} (2 \pi )^{9/2}}} &=& \eta_{q \bar q}\sqrt{1 \over 3} {{\tilde m_D \tilde m_{\pi}^{1/2} \beta^3_D \beta^{3/2}_{\pi} \beta^{3/2}_{ft}} \over {8 \pi^6}} \int d^3r \int d^3v \int d^3w \nonumber \\ && e^{-{1 \over 2} \beta^2_D(\vec w + {1 \over 2} \vec v)^2 -{1 \over 2} \beta^2_{\pi}(\vec r- \vec w + {1 \over 2} \vec v)^2 -{1 \over 2} \beta^2_{ft} w^2 -{1 \over 2} \beta^2_D r^2} {1 \over {(2 \pi)^{3/2}}}e^{-i{\vec q \over 2} \cdot (\vec r + \vec w - {1 \over 2} \vec v)} \psi_{pc}(\vec v)_+ \end{eqnarray} in which the ${1 \over \sqrt{2}}$ for the pure $\pi^0$ decay via $\eta_{u \bar u}$ has been explicitly included, $\tilde m_X$ is the ``mock meson" mass given by the sum of the constituent quark masses, and $\psi_{pc}(\vec v)_+={\beta^{5/2}_{pc} \over \pi^{3/4}}v_+e^{-{1 \over 2} \beta^2_{pc} v^2}$. The integrals are straightforward and give \begin{equation} g_0=-\eta_{q \bar q} { {16 \pi^{3/4} \tilde m_D \tilde m_{\pi}^{1/2} \beta^3_D \beta^{3/2}_{\pi} \beta^{3/2}_{ft} \beta^{5/2}_{pc}({1 \over 2}-a_{DD} -b_{DD})} \over {\sqrt{3}{\beta_v}_{DD}^5 {\beta_y}_{DD}^3 {\beta_w}_{DD}^3 } } e^{-{q^2 \over {8\bar \beta_{DD}^2}}} \end{equation} where \begin{eqnarray} {1 \over {\bar \beta_{DD}^2}}&=& {1 \over {\beta^2_y}}+ {1 \over {\beta^2_w}} \bigl( {{\beta^2_D+2 \beta^2_{\pi}} \over {\beta^2_D+\beta^2_{\pi}}} \bigr)^2+ {1 \over {\beta^2_v}}({1 \over 2}-a-b)^2\\ {\beta_v}_{DD}^5&=& \beta^2_{pc} +a^2(\beta^2_D+\beta^2_{\pi}+\beta^2_{ft}) +(b^2+{1 \over 4})(\beta^2_D+\beta^2_{\pi}) +a\beta^2_D+ (b-2ab-a)\beta^2_{\pi} \\ {\beta_y}_{DD}^2&=& \beta^2_D+\beta^2_{\pi} \label{eq:betay} \\ {\beta_w}_{DD}^2&=& \beta^2_D + \beta^2_{ft} +{{\beta^2_{\pi} \beta^2_D} \over {\beta^2_D+\beta^2_{\pi}}} \\ a_{DD}&=&- {\beta^4_D \over {2 [(\beta^2_D+\beta^2_{\pi}) (\beta^2_D+\beta^2_{\pi}+\beta^2_{ft}) - \beta^4_{\pi}]}} \\ b_{DD}&=&- {{\beta^2_{\pi}(2\beta^2_D+\beta^2_{ft}) } \over {2 [(\beta^2_D+\beta^2_{\pi}) (\beta^2_D+\beta^2_{\pi}+\beta^2_{ft}) - \beta^4_{\pi}]}}~~. \label{eq:b} \end{eqnarray} From the ``measured" $D^{0*} \rightarrow D^0 \pi^0$ width, we can deduce that $g_0 \simeq 11$. With our canonical parameters it follows that, with $\beta_{pc}$ expressed in GeV, \begin{equation} \eta_{q \bar q} \simeq 0.32[{{3+2f} \over f^{1/2}}]^{3/2}[{{\beta_{pc}^2 + 0.06} \over \beta_{pc}}]^{5/2} \end{equation} where since $\beta_D \simeq \beta_{\pi} \simeq \beta_{ft}/f^{1/2}$ I have been able to explicitly display the dependence on $f$ as well as $\beta_{pc}$. We see that for a variation of $\pm 1$ around the canonical value $f=2$, $\eta_{q \bar q}$ varies by less than $10 \%$. Thus we conclude that (since $\beta^2_{pc} >> 0.06$ GeV$^2$), \begin{equation} \eta_{q \bar q} \simeq 0.9 [{\beta_{pc} \over 0.58~ \rm{GeV}}]^{5/2}~ \rm{GeV} \end{equation} {\it i.e.}, $\eta_{q \bar q} \simeq 0.9$ for the canonical value $\beta_{pc}=0.58$ GeV. We also see from this formula the expected result that as the pair creation operator becomes more pointlike, $\eta_{q \bar q} \rightarrow \infty$. \vfill\eject {\noindent \bf APPENDIX C: Calculating $\tau^{(m)}_{1/2}$ and $\tau^{(p)}_{3/2}$ for Selected Low-lying Exclusive Nonresonant Channels} \bigskip As discussed in Ref. \cite{IWonBj}, the semileptonic decays $\bar B \rightarrow D^{(n)}_{{s'}_{\ell}^{{\pi '}_{\ell}}} \ell \bar \nu_{\ell}$ are governed in the heavy quark limit by generalized Isgur-Wise functions which determine all of the form factors for the decay of the $\bar B$ with $s_{\ell}^{\pi_{\ell}}={1\over 2}^-$ to both of the states of a heavy quark spin multiplet with quantum numbers ${s'}_{\ell}^{{\pi '}_{\ell}}$. As described in the text and elsewhere \cite{IWonBj}, \begin{equation} \rho^2={1\over 4}+\Delta \rho^2_{pert}+\rho^2_{Q\bar d}+\Delta \rho^2_{sea} \end{equation} where the ${1\over 4}$ is Bjorken's relativistic correction \cite{Bj}, $\Delta \rho^2_{pert}$ is a perturbative QCD radiative correction, and $\rho^2_{Q\bar d}$ and $\Delta \rho^2_{sea}$ are the contributions to the slope of $\xi(w)$ from the valence and sea quarks, respectively. As we have seen, these latter two contributions may be related to the rates of decay into inelastic channels by \begin{eqnarray} \rho^2_{Q\bar d}&=& \sum_{m={1\over 2}^+ Q \bar d~resonances}\vert \tau^{(m)}_{1/2} \vert^2 +2 \sum_{p={3\over 2}^+ Q \bar d~resonances}\vert \tau^{(p)}_{3/2} \vert^2 \\ \rho^2_{sea}&=& \sum_{m={1\over 2}^+ continuum}\vert \tau^{(m)}_{1/2} \vert^2 +2 \sum_{p={3\over 2}^+ continuum}\vert \tau^{(p)}_{3/2} \vert^2 \label{eq:tauplustau} \end{eqnarray} where the $\tau$'s are the appropriate Isgur-Wise functions. Once the $\tau$'s are specified, all transition form factors to the states of a heavy quark spin multiplet may be determined from symmetry considerations. For this reason, it is useful to calculate the $\tau$'s in the simplest possible setting, namely for the case where $b$ and $c$ are {\it spinless}. This is possible since the $\tau$'s depend only on the dynamics of the light degrees of freedom, {\it i.e.}, on the transition $s_{\ell}^{\pi_{\ell}}={1\over 2}^- \rightarrow {s'}_{\ell}^{{\pi '}_{\ell}}={1\over 2}^+$ or ${3\over 2}^+$. Another key simplification arises from the dynamics of the pair creation process. As demonstrated in the text, when $b(\vec v) \rightarrow c(\vec {v'})$ in a $Q \bar q q \bar d$ state, excitation of the variable $\vec r$ cannot lead to a contribution of order $(w-1)$. Furthermore, the variable $\vec v$ cannot be excited since this is a $q \bar q$ internal coordinate. Thus to contribute at order $(w-1)$, $b(\vec v) \rightarrow c(\vec {v'})$ {\it must} kick the $\vec w$ coordinate into an $\ell_w=1$ state: {\it such a state is the parent of all nonresonant production to order} $(w-1)$. Thus $\Delta \rho^2_{sea}$ arises entirely from the ``decay" of the lowest $\ell_w=1$ excited state of $c \bar q q \bar d$ arising from the $b \rightarrow c$ transition from $b \bar q q \bar d$. With the $q \bar q$ pair in $J^P=0^+$, the decay can thus occur from the six states $w_+ \uparrow$, $w_+ \downarrow$, $w_0 \uparrow$, $w_0 \downarrow$, $w_- \uparrow$, and $w_- \downarrow$, depending on which component of $\vec w$ is excited by the recoil. Here $w_+, w_0, w_-$ represent the components of the $\ell_w=1$ state, and $\uparrow, \downarrow$ represent the spin state of the $\bar d$ spectator quark. Since the total decay rate of $b \bar q q \bar d \uparrow$ and $b \bar q q \bar d \downarrow$ must be the same, we can simplify if we average rates over the initial $\bar d$ spin and over directions of $\vec P_{cm}$ (or equivalently, over directions of $\vec w$). Then since the average over the six states just listed will be the same as the average over the two $j={1\over 2}^+$ and four $j={3\over 2}^+$ states formed from them, we can deal with ``parent" states that are states with good angular momenta and which therefore uniquely feed the ${s'}_{\ell}^{{\pi '}_{\ell}}={1\over 2}^+$ and ${3\over 2}^+$ states, respectively. Let me provide an example: the production of the $(D+D^*)\pi$ nonresonant states. Since we know that all of $\Delta \rho^2_{sea}$ arises from the ``parent" state, the fraction of $\Delta \rho^2_{sea}$ coming from a given channel can just be obtained as the $jm_j$ average of the square of an overlap between a given $jm_j$ in $Q \bar q q \bar d$ and the two particle continuum state of interest. Thus from the $jm_j={1\over 2}{1\over 2}$ state we can extract $\langle (D+D^*)_{{1 \over 2}^- {1 \over 2}} \pi (\vec q~) \vert c \bar q q \bar d ;{1\over 2}^+{1\over 2} \rangle$, which ought to leave the $(D+D^*)\pi$ system in an $S$-wave. Explicitly, as $m_Q \rightarrow \infty$, \begin{eqnarray} \langle (D+D^*)_{{1 \over 2}^- {1 \over 2}} \pi (\vec q~) \vert c \bar q q \bar d ;{1\over 2}^+{1\over 2} \rangle &=& \int d^3w \int d^3v \int d^3r {\beta_D^{3/2} \over \pi^{3/4}}e^{-{1 \over 2}\beta_D^2(\vec w+ {1 \over 2}\vec v)^2} {\beta_{\pi}^{3/2} \over \pi^{3/4}}e^{-{1 \over 2}\beta_{\pi}^2(\vec r-\vec w+ {1 \over 2}\vec v)^2} \nonumber \\ && {1 \over {(2\pi)^{3/2}}}e^{-i{\vec q \over 2} \cdot (\vec r+ \vec w - {1 \over 2}\vec v)} {\tilde \beta_{pc}^{5/2} \over \pi^{3/4}}e^{-{1 \over 2}\beta_{pc}^2 v^2} {\beta_{ft}^{5/2} \over \pi^{3/4}}e^{-{1 \over 2}\beta_{ft}^2 w^2} {\beta_B^{3/2} \over \pi^{3/4}}e^{-{1 \over 2}\beta_B^2 r^2} \nonumber \\ && \nonumber \\ && \Sigma (\vec w, \vec v) \end{eqnarray} where \begin{equation} \Sigma \equiv \langle \uparrow \sqrt{1 \over 2} (\uparrow \downarrow - \downarrow \uparrow) \vert \sqrt{1 \over 3} (\uparrow \uparrow v_--[\uparrow \downarrow + \downarrow \uparrow]v_z- \downarrow \downarrow v_+) \sqrt{2 \over 3}(-w_+\downarrow-w_z\uparrow) \rangle \label{eq:Dpiintegral} \end{equation} is the spin overlap matrix element of the three quarks $\bar q q \bar d$, respectively. We get \begin{equation} \langle (D+D^*)_{{1 \over 2}^- {1 \over 2}} \pi (\vec q~) \vert c \bar q q \bar d ;{1\over 2}^+{1\over 2} \rangle =-I^{00}_{+-} - I^{00}_{zz} \end{equation} where \begin{equation} I^{00}_{ij} \equiv {1 \over 3} { {\beta_D^{3/2}\beta_{\pi}^{3/2}\beta_{ft}^{5/2}\tilde \beta_{pc}^{5/2}\beta_B^{3/2}} \over {\pi^{15/4}} } \int d^3w \int d^3v \int d^3r {{v_iw_j} \over {(2\pi)^{3/2}}}e^{-i{\vec q \over 2} \cdot (\vec r+ \vec w - {1 \over 2}\vec v)-{1 \over 2}E^2} \end{equation} with \begin{equation} E^2= \beta_D^2 (\vec w + {1 \over 2}\vec v)^2 +\beta_{\pi}^2 (\vec r - \vec w + {1 \over 2}\vec v)^2 +\beta_{ft}^2 w^2 +\tilde \beta_{pc}^2 v^2 +\beta_B^2 r^2~. \end{equation} In these formulas I have distinguished between $\beta_D^2$ and $\beta_B^2$ to allow the ``ancestry" of terms to be traced, even though $\beta_B=\beta_D$ from heavy quark symmetry. This integral is easily done, giving \begin{equation} I^{00}_{ij}=I^{00} \Bigl[ c^{00} \delta_{ij} + c^{00}_{ij} {{q_iq_j} \over \beta^2_v} \Bigr] e^{-q^2/ 8\tilde{\bar \beta}^2} \end{equation} where $I^{00}$, $c^{00}_{ij}$, $c^{00}$, and $\tilde{\bar \beta}$ are given below. For the problem at hand, we get \begin{eqnarray} \langle (D+D^*)_{{1 \over 2}^- {1 \over 2}} \pi (\vec q~) \vert c \bar q q \bar d ;{1\over 2}^+{1\over 2} \rangle &=&-\Bigl[3c^{00}+ c^{00}_{ij}{{q_+q_-} \over \beta^2_v} +c^{00}_{ij}{q^2_z \over \beta^2_v} \Bigr] I^{00}e^{-q^2/ 8\tilde{\bar \beta}^2}\nonumber \\ &=&-\Bigl[3c^{00} + c^{00}_{ij} {{q^2} \over \beta^2_v} \Bigr] I^{00} e^{-q^2/ 8\tilde{\bar \beta}^2} \nonumber \\ &=&-\sqrt{4 \pi} \Bigl[3c^{00} + c^{00}_{ij} {{q^2} \over \beta^2_v} \Bigr] I^{00} e^{-q^2/ 8\tilde{\bar \beta}^2}Y_{00}(\Omega_q)~, \end{eqnarray} a pure $S$-wave decay as required, with partial wave amplitude $A_{1 / 2} \equiv -\sqrt{4 \pi} \Bigl[3c^{00} + c^{00}_{ij} {{q^2} \over \beta^2_v} \Bigr] I^{00} e^{-q^2/ 8\tilde{\bar \beta}^2}$. Note that with $Y_{00}$ factored out, $\vert A_{1 / 2} \vert^2$ is already the probability for this channel integrated over angles $\Omega_q$, leaving only an integral $\int dq q^2 \vert A_{1 / 2} \vert^2$ to be done to sum over all $(D+D^*) \pi $ states with quantum numbers ${1 \over 2}^+ {1 \over 2}$ at any fixed value of $(w-1)$. Next consider $\langle (D+D^*)_{{1 \over 2}^- {1 \over 2}} \pi (\vec q~) \vert c \bar q q \bar d ;{3\over 2}^+{3\over 2} \rangle$, which proceeds by replacing $\sqrt{2 \over 3}(-w_+\downarrow-w_z\uparrow) \rangle$ by $-w_+ \uparrow$ in Eq. (\ref{eq:Dpiintegral}). With this change one gets \begin{eqnarray} \langle (D+D^*)_{{1 \over 2}^- {1 \over 2}} \pi (\vec q~) \vert c \bar q q \bar d ;{3\over 2}^+{3\over 2} \rangle &=&-\sqrt{3 \over 2} I^{00}_{z+} \nonumber \\ &=&-\sqrt{3 \over 2} c^{00}_{ij} {q_zq_+ \over \beta^2_v}I^{00} e^{-q^2/ 8\tilde{\bar \beta}^2} \nonumber \\ &=&-\sqrt{1 \over 5}[\sqrt{4 \pi} c^{00}_{ij} {{q^2} \over \beta^2_v} I^{00}e^{-q^2/ 8\tilde{\bar \beta}^2}] Y_{21}(\Omega_q)~, \end{eqnarray} a pure $D$-wave decay as required. Moreover, since $-\sqrt{1 \over 5}$ is the Clebsch-Gordan coefficient for coupling $(D+D^*)_{{1 \over 2}^- {1 \over 2}}$ and an $\ell=2$, $m=1$ pion into a ${3 \over 2}{3 \over 2}$ state, we can deduce that decays to this whole angular momentum multiplet with $(D+D^*) \pi $ in a $D$-wave coupled to ${s'_{\ell}}^{{\pi '}_{\ell}}={3 \over 2}^+$ are controlled by a $D$-wave amplitude $A_{3 / 2} \equiv -\sqrt{4 \pi} c^{00}_{ij} {{q^2} \over \beta^2_v} I^{00}e^{-q^2/ 8\tilde{\bar \beta}^2}$. To complete this pedagogical example, I note that since the partial wave decay amplitudes are independent of the magnetic substate $m$, \begin{eqnarray} {1 \over 6} \sum_m \vert A_{{1 / 2}} \vert^2 &=& {1 \over 3}\vert A_{{1 / 2}} \vert^2 \label{eq:A1/2rate} \\ {1 \over 6} \sum_m \vert A_{{3 / 2}} \vert^2 &=& {2 \over 3}\vert A_{{3 / 2}} \vert^2 ~~, \label{eq:A3/2rate} \end{eqnarray} where the ${1 \over 6}$ arises from averaging over the six $jm_j$ states. On comparison with Eq. (\ref{eq:tauplustau}), we see that $\tau_{1 / 2}=\sqrt{1 \over 3} A_{{1 / 2}} \sqrt{{1 \over 3} \Delta \rho^2_{sea} }$ and $\tau_{3 / 2}=\sqrt{1 \over 3} A_{{3 / 2}} \sqrt{ {1 \over 3} \Delta \rho^2_{sea} }$. Note that ${1 \over 3} \Delta \rho^2_{sea}$ appears since the overlap amplitudes $A_{1 / 2}$ and $A_{3 / 2}$ as calculated are for a single flavor, while the factor $\sqrt{1 \over 3}$ arises as a residue of the angular and spin averaging. The tricks outlined here are more powerful for more complex decays. I will give one partial illustration: $\langle [(D+D^*) \rho]_{{3\over 2}^+{3\over 2}} \vert c \bar q q \bar d ;{3\over 2}^+{3\over 2} \rangle$, where the subscripts on the bracket are total spin quantum numbers, but do not include relative orbital angular momentum. The overlap integral for this decay is obtained by replacing the pion spin wavefunction $\sqrt{1 \over 2}(\uparrow \downarrow- \downarrow \uparrow)$ by $\uparrow \uparrow$, $\sqrt{2 \over 3}(-w_+ \downarrow-w_z \uparrow)$ by $-w_+ \uparrow$, and $\beta_{\pi}$ by $\beta_{\rho}$ in Eq. (\ref{eq:Dpiintegral}) to give \begin{eqnarray} \langle [(D+D^*) \rho]_{{3\over 2}^+{3\over 2}} \vert c \bar q q \bar d ;{3\over 2}^+{3\over 2} \rangle &=&-\sqrt{3} I^{00}_{+-} \nonumber \\ &=&-\sqrt{3} [ 2 c^{00} + c^{00}_{ij} {q_+q_- \over \beta^2_v}]I^{00} e^{-q^2/ 8\tilde{\bar \beta}^2} \nonumber \\ &=&-\sqrt{4 \over 3}\sqrt{4 \pi} \Bigl([ 3 c^{00} + c^{00}_{ij} {q^2 \over \beta^2_v} ] Y_{00} \nonumber \\ &&~~~~~~~- c^{00}_{ij} {q^2 \over \beta^2_v} \sqrt{1 \over 5} Y_{20} \Bigr)I^{00}e^{-q^2/ 8\tilde{\bar \beta}^2} ~, \end{eqnarray} By examining the Clebsch-Gordan coefficients for coupling the spin state $[(D+D^*) \rho]_{{3\over 2}^+{3\over 2}}$ to a relative orbital $S$-wave or a $D$-wave to get a ${3 \over 2}{3 \over 2}$ state, we can deduce that the $S$- and $D$-wave amplitudes for this decay are $- \sqrt{4 \over 3}[\sqrt{4 \pi} (3 c^{00} + c^{00}_{ij} {q^2 \over \beta^2_v} )I^{00}e^{-q^2/ 8\tilde{\bar \beta}^2}]$ and $\sqrt{4 \over 3}[\sqrt{4 \pi} c^{00}_{ij} {q^2 \over \beta^2_v} I^{00}e^{-q^2/ 8\tilde{\bar \beta}^2}]$, respectively. One very simple overlap integral thus gives two partial wave amplitudes simultaneously. A complete set of results are given in the text in Table II in terms of the following basic integrals: \begin{eqnarray} I^{00}_{ij} &=& I^{00} \Bigl[c^{00} \delta_{ij} + c^{00}_{ij} {{q_iq_j} \over \tilde \beta^2_v} \Bigr] e^{-{q^2 / {8 \tilde {\bar \beta}^2}}} \\ I^{10}_{ijk} &=& I^{10} \Bigl[c^{10}_i q_i \delta_{jk} + c^{10}_j q_j \delta_{ik} + c^{10}_k q_k \delta_{ij} + c^{10}_{ijk} {{q_i q_j q_k} \over {\tilde \beta^2_{v~**}}} \Bigr] e^{-{q^2 / {8 \tilde {\bar \beta}_{**}^2}}} \\ I^{01}_{ijk} &=& I^{01} \Bigl[c^{01}_i q_i \delta_{jk} + c^{01}_j q_j \delta_{ik} + c^{01}_k q_k \delta_{ij} + c^{01}_{ijk} {{q_i q_j q_k} \over {\tilde \beta^2_{v~a}}} \Bigr] e^{-{q^2 / {8 \tilde {\bar \beta}_a^2}}} \end{eqnarray} where $I^{00}_{ij}$ is specified in terms of \begin{eqnarray} I^{00}&=& { {2 \beta^{3/2}_D \beta^{3/2}_B \beta^{3/2}_{\pi} \beta^{5/2}_{ft} \tilde \beta^{5/2}_{pc}} \over {3 \pi^{3/4} \tilde \beta^5_v \beta^5_w \beta^3_y} } \\ c^{00}&=& +4a\beta^2_w \\ c^{00}_{ij}&=&\Bigl[({1 \over 2}-a -b) \bigl( {{\beta^2_B+2 \beta^2_{\pi}} \over {\beta^2_B+\beta^2_{\pi}}} \bigr) -{a \beta^2_w \over \tilde \beta^2_v}({1 \over 2}-a -b)^2\Bigr] \\ \tilde \beta^2_v&=& \tilde \beta^2_{pc} +a^2(\beta^2_D+\beta^2_{\pi}+\beta^2_{ft}) +b^2(\beta^2_B+\beta^2_{\pi}) \nonumber \\ &&+{1 \over 4}(\beta^2_D+\beta^2_{\pi}) +a\beta^2_D+ (b-2ab-a)\beta^2_{\pi} \\ \beta_y^2 &=& \beta^2_B+\beta^2_{\pi} \\ \beta_w^2 &=& \beta^2_D + \beta^2_{ft} +{{ \beta^2_B \beta^2_{\pi} } \over {\beta^2_B+\beta^2_{\pi}}} \\ a &=&- {{(\beta^2_B+\beta^2_{\pi})(\beta^2_D-\beta^2_{\pi})+\pi^4} \over {2 [(\beta^2_B+\beta^2_{\pi}) (\beta^2_D+\beta^2_{\pi}+\beta^2_{ft}) - \beta^4_{\pi}]}} \\ b&=&- {{\beta^2_{\pi}(2\beta^2_D+\beta^2_{ft}) } \over {2 [(\beta^2_B+\beta^2_{\pi}) (\beta^2_D+\beta^2_{\pi}+\beta^2_{ft}) - \beta^4_{\pi}]}} \\ {1 \over { \tilde {\bar \beta}^2}}&=& {1 \over {\beta^2_y}}+ {1 \over {\beta^2_w}} \bigl( {{\beta^2_B+2 \beta^2_{\pi}} \over {\beta^2_B+\beta^2_{\pi}}} \bigr)^2+ {1 \over {\tilde \beta^2_v}}({1 \over 2}-a-b)^2~~. \end{eqnarray} Similarly, $I^{10}_{ijk}$ is specified in terms of \begin{eqnarray} I^{10}&=& - { {4i \beta^{5/2}_{D^{**}} \beta^{3/2}_B \beta^{3/2}_{\pi} \beta^{5/2}_{ft} \tilde \beta^{5/2}_{pc}} \over {3 \pi^{3/4} \tilde {\beta_v}_{**}^5 {\beta_w}_{**}^5 {\beta_y}_{**}^3} } \\ c^{10}_{i}&=&\Bigl[+a_{**}\bigl( {{\beta^2_B+2 \beta^2_{\pi}} \over {\beta^2_B+\beta^2_{\pi}}} \bigr) -a_{**} ({1 \over 2}+a_{**}) ({1 \over 2}-a_{**} -b_{**}) {{\beta_w}_{**}^2 \over \tilde {\beta_v}_{**}^2} \Bigr] \\ c^{10}_{j}&=&\Bigl[-({1 \over 2}-a_{**} -b_{**}) -a_{**} ({1 \over 2}+a_{**}) ({1 \over 2}-a_{**} -b_{**}) {{\beta_w}_{**}^2 \over \tilde {\beta_v}_{**}^2} \Bigr] \\ c^{10}_{k}&=&\Bigl[+({1 \over 2}+a_{**}) \bigl( {{\beta^2_B+2 \beta^2_{\pi}} \over {\beta^2_B+\beta^2_{\pi}}} \bigr) -a_{**} ({1 \over 2}+a_{**}) ({1 \over 2}-a_{**} -b_{**}) {{\beta_w}_{**}^2 \over \tilde {\beta_v}_{**}^2} \Bigr] \\ c^{10}_{ijk}&=&+{1 \over 4} \Bigl[({1 \over 2}-a_{**} -b_{**}) {\tilde {\beta_v}_{**}^2 \over {\beta_w}_{**}^2} \bigl( {{\beta^2_B+2 \beta^2_{\pi}} \over {\beta^2_B+\beta^2_{\pi}}} \bigr)^2 \nonumber \\ && -({1 \over 2}+2a_{**})({1 \over 2}-a_{**} -b_{**})^2 \bigl( {{\beta^2_B+2 \beta^2_{\pi}} \over {\beta^2_B+\beta^2_{\pi}}} \bigr) \nonumber \\ &&+a_{**} ({1 \over 2}+a_{**}) ({1 \over 2}-a_{**} -b_{**})^3 {{\beta_w}_{**}^2 \over \tilde {\beta_v}_{**}^2} \Bigr] \end{eqnarray} where $\tilde {\beta_v}_{**}^2$, ${\beta_y}_{**}^2$, ${\beta_w}_{**}^2$, $a_{**}$, $b_{**}$, and ${1 \over { \tilde {\bar \beta}_{**}^2}}$ are given by the formulas for the $I^{00}_{ij}$ variables $\tilde \beta^2_v$, $\beta^2_y$, $\beta^2_w$, $a$, $b$, and ${1 \over { \tilde {\bar \beta}^2}}$, respectively, with $\beta_D \rightarrow \beta_{D^{**}}$ everywhere. Finally, $I^{01}_{ijk}$ is specified in terms of \begin{eqnarray} I^{01}&=& - { {4i \beta^{3/2}_D \beta^{3/2}_B \beta^{5/2}_a \beta^{5/2}_{ft} \tilde \beta^{5/2}_{pc}} \over {3 \pi^{3/4} \tilde {\beta_v}_a^5 {\beta_w}_a^5 {\beta_y}_a^3} } \\ c^{01}_{i}&=&\Bigl[+ {a_a {\beta_w}_a^2 \over \tilde {\beta_v}_a^2} -{{a_a \beta^2_B} \over {{\beta^2_B+\beta^2_a}}} \bigl({{\beta^2_B+2 \beta^2_a} \over {\beta^2_B+\beta^2_a}} \bigr) -a_a ({1 \over 2}-a_a+b_a) ({1 \over 2}-a_a -b_a) {{\beta_w}_a^2 \over \tilde {\beta_v}_a^2} \Bigr] \\ c^{01}_{j}&=&\Bigl[+{{\beta^2_B} \over {{\beta^2_B+\beta^2_a}}} ({1 \over 2}-a_a -b_a) -a_a ({1 \over 2}-a_a+b_a) ({1 \over 2}-a_a -b_a) {{\beta_w}_a^2 \over \tilde {\beta_v}_a^2} \Bigr] \\ c^{01}_{k}&=&\Bigl[+\bigl({{\beta^2_B+2 \beta^2_a} \over {\beta^2_B+\beta^2_a}} \bigr)({1 \over 2}-a_a+b_a) -a_a ({1 \over 2}-a_a+b_a) ({1 \over 2}-a_a -b_a) {{\beta_w}_a^2 \over \tilde {\beta_v}_a^2} \Bigr] \\ c^{01}_{ijk}&=&+{1 \over 4} \Bigl[ +({1 \over 2}-a_a -b_a) {\tilde {\beta_v}_a^2 \over {\beta_y}_a^2} \bigl( {{\beta^2_B+2 \beta^2_a} \over {\beta^2_B+\beta^2_a}} \bigr) -{a_a {\beta_w}_a^2 \over {\beta_y}_a^2}({1 \over 2}-a_a -b_a)^2 \nonumber \\ && -{\tilde {\beta_v}_a^2 \over {\beta_w}_a^2} \bigl( {{\beta^2_B+2 \beta^2_a} \over {\beta^2_B+\beta^2_a}} \bigr)^2 {{\beta^2_B} \over {{\beta^2_B+\beta^2_a}}} ({1 \over 2}-a_a -b_a) \nonumber \\ && +{{ a_a \beta^2_B} \over {{\beta^2_B+\beta^2_a}}} \bigl( {{\beta^2_B+2 \beta^2_a} \over {\beta^2_B+\beta^2_a}} \bigr) ({1 \over 2}-a_a-b_a)^2 \nonumber \\ && -\bigl( {{\beta^2_B+2 \beta^2_a} \over {\beta^2_B+\beta^2_a}} \bigr) ({1 \over 2}-a_a+b_a)({1 \over 2}-a_a -b_a)^2 \nonumber \\ && +{a_a {\beta_w}_a^2 \over \tilde {\beta_v}_a^2}({1 \over 2}-a_a+b_a)({1 \over 2}-a_a -b_a)^3 \Bigr] \end{eqnarray} where $\tilde {\beta_v}_a^2$, ${\beta_y}_a^2$, ${\beta_w}_a^2$, $a_a$, $b_a$, and ${1 \over { \tilde {\bar \beta}_a^2}}$ are given by the formulas for the $I^{00}_{ij}$ variables $\tilde \beta^2_v$, $\beta^2_y$, $\beta^2_w$, $a$, $b$, and ${1 \over { \tilde {\bar \beta}^2}}$, respectively, with $\beta_{\pi} \rightarrow \beta_a$ everywhere. \vfill\eject {\centerline {\bf REFERENCES}}
\section{Dark Matter and Cosmology} The two pillars of standard cosmology are the Einstein equations \beq {\rm R}_{\mu\nu} - {1\over 2} {\cal R} {\rm g}_{\mu\nu} = 8 \pi {\rm G} {\rm T}_{\mu\nu} + {\rm \Lambda} {\rm g}_{\mu\nu} \, , \label{eq:einstein} \eeq \noindent (where ${\rm R}_{\mu \nu}$ is the curvature (Ricci) tensor, $\cal R$ is its trace, ${\rm T}_{\mu \nu}$ is the stress-energy tensor, ${\rm \Lambda}$ the cosmological constant) and the Robertson-Walker metric \beq ds^2 = dt^2 - {\rm R}^2(t) \left({ \frac{dr^2}{1-k r^2} + r^2 d{\theta}^2 + r^2 \sin^2 \theta \, d{\phi}^2} \right) \, , \label{eq:rw} \eeq \noindent where $R = R(t)$ is the cosmic scale factor and $t,r,\theta,\phi$ are the comoving coordinates \cite{weinberg,kt,peebles,roos,kolb}. We recall that Eq. (\ref{eq:rw}) follows from the assumption that the distribution of matter and radiation in the Universe is isotropic and homogeneous. When the coordinates are appropriately rescaled, the values $k = +1, 0, -1$ define the space curvature to be positive, zero and negative, respectively. Useful standard quantities for the description of the expanding Universe are the Hubble parameter (expansion rate of the Universe) $H = \dot{\rm R}/R$ and the deceleration parameter $q = - \ddot{\rm R} \cdot {\rm R}$/$\dot{\rm R}^2$, whose values at present epoch are denoted as ${\rm H}_0$ and ${\rm q}_0$. By combining Eqs. (\ref{eq:einstein}) -- (\ref{eq:rw}), one obtains the Friedmann equation \beq {\rm \Omega}_m + {\rm \Omega}_{\rm \Lambda} - k/({\rm H}^2 {\rm R}^2) = 1\, , \label{eq:friedmann} \eeq \noindent where ${\rm \Omega}_m$ is the ratio of the average matter--energy density $\rho$ to the critical density $\rho_c = 3 {\rm H}^2/(8 \pi G)$, $\Omega_m = \rho/\rho_c$, and ${\rm \Omega}_{\rm \Lambda} = {\rm \Lambda} /(3 {\rm H}^2)$. The critical density is given by $\rho_c = 1.9 \cdot 10^{-29}\, h^2$ g\ cm$^{-3}$, when the Hubble constant is parametrised as follows: $h = {\rm H}_0/(100$ km\ s$^{-1}$ Mpc$^{-1}$). {}From Eqs. (\ref{eq:einstein}) -- (\ref{eq:rw}) it also follows that at present time, when the radiation contribution to $\Omega$ can be set to zero, the value of the deceleration parameter is given by \beq q_0 = \frac{1} {2} {\Omega}_m - {\Omega}_{\Lambda}\, . \label{eq:q} \eeq The features of the evolution of our Universe depend on the actual values to be assigned to the cosmological parameters previously defined. In what follows we briefly summarize some of the main observational data about these parameters. \section{Age of the Universe and Expansion Rate} The evaluation of the present age of the Universe, $t_0$, depends on the expansion rate and on the various components of $\Omega$; therefore measurements of $t_0$ and of the Hubble constant provide information on the size of $\Omega_m$ and $\Omega_{\Lambda}$ (see, for instance, Refs. \cite{kt,freedman}). A recent determination of $t_0$ provides the value: $t_0 = 11.5 \pm 1.3$ Gyr \cite{cdkk}, with a $95\%$ C.L. lower bound of 9.5 Gyr. Recent measurements of the Hubble constant by the Hubble Space Telescope Key project give the following range \cite{freedman,madore}: ${\rm H}_0 = 73 \pm 6 (stat) \pm 8 (syst)$ km s$^{-1}$ Mpc$^{-1}$. In view of the still persisting sizeable spread in the $h$ values, due to a host of independent measurements (for a review, see for instance \cite{freedman}) in the following we will, conservatively, consider a rather wide range: $0.5 \leq h \leq 0.9$. If, for sake of illustration, we take $h \simeq 0.7$, it is easy to show that $t_0 \sim 11.5$ Gyr would require either $\Omega_m \lsim 0.3$ or $\Omega_m \sim 0.5$, according to whether or not we allow for a non-vanishing cosmological constant, such that $\Omega_{\Lambda} = 1 - \Omega_m$. \section{Observational Evidence for Dark Matter} The parameter $\Omega_m$ may be derived from astronomical determinations of the average mass-to-light ratio $M/L$ for various astrophysical objects (see, for instance, Ref. \cite{sap} for an updated review) \beq \Omega_m = \frac{\rm M}{\rm L} \frac{\it L}{\rho_c}\, , \label{eq:ml} \eeq \noindent where $L$ is the luminosity density: ${\it L} = 1.6 \cdot 10^8\, h\, {\rm L}_{\odot}$ Mpc$^{-3}$. Using $\rho_c = 2.8 \cdot 10^{11}\, h^2\, {\rm M}_{\odot}$ Mpc$^{-3}$, one gets \beq \Omega_m = \frac{h^{-1}}{1500}\, \frac{{\rm M}/{\rm L}} {{\rm M}_{\odot}/{\rm L}_{\odot}} \, . \label{eq:aa} \eeq {}From the $M/L$ ratio measured in stars, $M/L \sim (3-9) {\rm M}_{\odot}/{\rm L}_{\odot}$ we obtain \beq 0.002 \;h^{-1} \leq \Omega_{vis} \leq 0.006 \; h^{-1} \, . \label{eq:bb} \eeq \subsection{Rotational curves of spiral galaxies} Presence of dark matter in single galaxies is apparent from the flatness of the plot of the rotational velocities versus the galactocentric radius, well beyond the distribution of the visible matter. {}From these measurements one derives $M/L \simeq 70 \, {\rm M}_{\odot}/{\rm L}_{\odot} (R_{halo}/100$ kpc). Then we obtain \beq \Omega_{halo} \sim 0.05\, h\, R_{halo}/100\, \mbox{{\rm kpc}}\, , \eeq \noindent a result which, compared to the range (\ref{eq:bb}) for $\Omega_{vis}$, is indicative of the presence of dark matter at the level of single galaxies. \subsection{Clusters of galaxies} Existence of dark matter at the level of clusters of galaxies may be established by a number of methods: X-ray emission by hot gas in intra cluster plasma, measurements of velocity dispersion and strong gravitational lensing. We report here some results derived from measurements of the X-ray emission \cite{wf}. For rich clusters of galaxies one finds $M/L = (300 \pm 100)\,h \,{\rm M}_{\odot}/{\rm L}_{\odot}$, which gives \beq \Omega_{cluster} \simeq 0.2 \pm 0.07 \, . \label{eq:cc} \eeq \noindent The baryonic content $\rm \Omega_b$ is established to be: $\sim 6\;h^{-3/2}\%$ for gas and $ \gsim 4\%$ for stars; then \beq \frac{\Omega_b}{\Omega_{cluster}} \gsim 0.04 + 0.06 h^{-3/2} \, . \label{eq:dd} \eeq On the other hand, the Big Bang nucleosynthesis provides the following estimate for $\Omega_b$ \cite{sap,nucl}: \begin{equation} 0.009 \lsim \Omega_b h^2 \lsim 0.02 \, , \label{eq:ob} \end{equation} or, taking $0.5 \leq h \leq 0.9$, $0.01 \lsim {\rm \Omega_b} \lsim 0.08$. Combining this result with Eq.(\ref{eq:dd}) we obtain \beq {\Omega}_{cluster} \lsim 0.2 - 0.4 \, . \label{eq:ee} \eeq \subsection{Large-scale Velocity Flows} Distribution of matter over large scales may be inferred from the peculiar motion of the gravitationally--induced inflow. Let us consider the case of the Local Supercluster, centered near the Virgo cluster; we stay on the edge of this cluster, at a distance of R = 12 $h^{-1}$ Mpc. The radial inward peculiar velocity averaged over the surface is given by \cite{peebles2} \beq \bar v = \frac{1}{3} {\rm H}_0 \, {\rm R} \delta_m {\Omega}^{0.6} \, . \label{eq:xx} \eeq \noindent where ${\Omega}^{0.6}$ represents the factor due to expansion and $\delta_m$ denotes the mass contrast, related to the contrast in galaxy counts ($\delta_g$) by the relation $\delta_m = \delta_g/b$, $b$ being the bias factor. Using the observational values \cite{peebles2}: $\bar v = 200 \pm 25$ km s$^{-1}$, $\delta_g = 2.3 \pm 0.7$, and taking $b \simeq 1$, one obtains \beq {\Omega}_m \simeq 0.1_{-0.05}^{+0.1}\, . \eeq \subsection{A few first conclusions} {}From the previous evaluations for $\Omega_{vis}$ and $\Omega_m$, and from the range for $\rm \Omega_b$ in Eq.(\ref{eq:ob}) we derive the following important points: \begin{itemize} \item a large amount of matter in the Universe is not visible; \item some of this dark matter is baryonic; \item a significant amount of dark matter is \underline{non}-- baryonic. \end{itemize} As usual we divide particle dark matter into two categories: Hot Dark Matter (HDM) and Cold Dark Matter (CDM), according to whether the particles are relativistic or non-relativistic at their decoupling from the primeval plasma. \section{Singling out different contributions to $\rm \Omega$} We briefly report now some important observational results and analyses which provide further clues toward a determination of the various components to $\rm \Omega$. The two main issues are: what, if any, is the size of $\rm \Omega_{\Lambda}$ and what is $\Omega_m$ made of? \subsection{Formation of cosmological structures} A standard method for testing different dark matter models is to compare the power spectrum of the density fluctuations $P(k)$ with measurements of the Cosmic Microwave Background Radiation (CMBR) anisotropy and measurements of the two-point correlation function in galaxy surveys \cite{kt,peebles,roos,kolb}. We recall that $P(k)$ is the Fourier transform of the correlation function between the density contrasts at two different points in space. It is customary to assume for $P(k)$ a power law, i.e. $P(k) \propto k^n$. The Harrison--Zel'dovich spectrum corresponds to $n = 1$. In the past, typically, the best performing cosmological models turned out to fall into the following categories: \begin{itemize} \item HCDM model characterized by $\rm \Omega = 1$ and the following composition: $\rm \Omega_b \simeq 0.1$, $\rm \Omega_{\nu} \simeq 0.2$, $\rm \Omega_{CDM} \simeq 0.7$; $h \simeq 0.5$, where $\rm \Omega_b$, $\rm \Omega_{\nu}$ and $\rm \Omega_{CDM}$ are the contribution due to baryons, neutrinos (as HDM particles) and to CDM particles, respectively. \item $\rm \Lambda$CDM model with $\rm \Omega = 1$ and $\rm \Omega_{CDM} \simeq 0.3$, $\rm \Omega_{\rm \Lambda} \simeq 0.7$; $h \simeq $ 0.7 -- 0.8. \item TCDM model ($\equiv$ tilted CDM model) with a power spectrum $P(k) \propto k^{0.8}$ and $\rm \Omega = 1 = \rm \Omega_{CDM} = 1$; $h \simeq 0.5$. \end{itemize} \subsection{Evidence for $\rm \Omega_{\rm \Lambda} \neq 0$ ?} Recent measurements of high-redshift supernovae of type Ia \cite{perl1,perl2,reiss} point to an important contribution to $\rm \Omega$ due to $\rm \Lambda$, with a relatively small contribution from $\Omega_m$. These data appear to be complementary to those derived from measurements of the CMBR \cite{new}. The joint use of these two sets of data, together with some other observational data, singles out the following ranges: $\Omega_m = 0.24 \pm 0.10$ and $\Omega_{\Lambda} = 0.62 \pm 0.16$ \cite{lineweaver}. These data and analyses are of the utmost interest for their potential implications, although a number of points need further clarification \cite{peebles}. One further hint for a rather low value of $\rm \Omega_m$ is provided by the time evolution of the number density of clusters. In Ref.\cite{neta} observational data on cluster abundance in the redshift range $0 \lsim z \lsim 1$ is used to derive the estimate $\rm \Omega_m \simeq 0.2^{+0.15}_{-0.1}$. Though a cautionary attitude is in order here, it is important to remark that all these different ways of determining $\rm \Omega_m$ point to a relative small value for this quantity: $\rm \Omega_m \lsim$ 0.3 -- 0.4. This feature, if confirmed by further observational data, has profound implications for the phenomenology related to DM particle candidates, as will be illustrated in the following. \section{Particle candidates for dark matter} \subsection{Baryons} As was already noticed in Sect. III, some contribution to DM is provided by baryons. This conclusion is drawn from the fact that the Big Bang nucleosynthesis provides the estimate $0.009 \lsim \Omega_b h^2 \lsim 0.02$, which, together with Eq. (\ref{eq:bb}), implies ${\rm \Omega_b} > {\rm \Omega_{vis}}$. Direct search for non-luminous baryonic dark matter, under the form of microlensing objects, has been undertaken since the seminal paper of Ref.\cite{pac}. Recent results are reviewed in \cite{sap,spiro,stubbs}. The main features of the data may be summarized as follows. The EROS Collaboration \cite{eros} excludes that microlensing objects of masses in the range $(5\cdot10^{-8} - 10^{-2})\, M_{\odot}$ may make up more than 20\% of the halo density, whereas the MACHO Collaboration \cite{macho} delimits a likelihood contour (at 95\% C.L.) for masses $\sim (10^{-1} - 1) \, M_{\odot}$ with a best-fit value for the halo fraction around 0.5. For further details see, for instance, Refs. \cite{sap,spiro,stubbs}. For a critical view of the microlensing events, see Ref.\cite{FFG}. An interesting scenario related to baryonic dark matter is the one depicted in Ref.\cite{jetzer}. \subsection{Non-baryonic DM candidates} Particle physics offers a large variety of particles, which would have decoupled from the primeval plasma at the time (freeze--out time), when the interaction rate became smaller than the cosmic expansion rate; these particles would then be floating around in our Universe as relics. These fossil particles would or would not significantly contribute to the average cosmic density depending on their actual number density and mass. The most obvious example is provided by light fossil neutrinos, whose relic abundance may be easily evaluated (see, for instance Ref.\cite{kt}) and turns out to be \beq {\rm \Omega}_{\nu} h^2 = \frac {\sum m_{\nu}} {93 \, \mbox{\rm eV}}\, , \label{eq:neu} \eeq \noindent where the sum is over the neutrino flavours (for each flavour, neutrino and antineutrino are counted together). Therefore the relevance of these fossil neutrinos for the Universe matter density depends on whether their mass is of order of a few eV or much smaller (provided neutrinos are massive at all). Possible indications for non--vanishing neutrino masses are from: 1) solar neutrinos \cite{solar,smy}, 2) atmospheric neutrinos \cite{sk,macro,habig}, and 3) the LSND experiment \cite{lsnd}. All these data refer to oscillation measurements, and then are not sensitive to individual neutrino masses, but only to differences in their squares. Typically, atmospheric neutrino experiments give $\Delta m^2 \simeq (2 \div 6) \times 10^{-3} \mbox{{eV}}^2$; solar neutrino experiments can be explained either by a $\Delta m^2 \simeq 10^{-5} \mbox{{eV}}^2$, in case of matter--enhanced oscillations, or by a $\Delta m^2 \simeq 10^{-10} \mbox{{eV}}^2$, in case of vacuum oscillations; LSND data suggest $0.2 \,\mbox{{eV}}^2 \lsim \Delta m^2 \lsim 2 \,\mbox{{eV}}^2$. These results may be compatible with a sizeable value of relic abundance, $\rm \Omega_{\nu} \sim 0.2$, such as preferred by some calculations of cosmological structures. However, if taken at their face values, they only imply ${\rm \Omega_{\nu}} \gsim 0.02 \, (0.5/h)^2$ \cite {prim}. A different candidate is the axion, whose motivation in particle physics is related to the strong CP-problem in QCD \cite{peccei}. A discussion of this candidate is beyond the scope of the present report; for a comprehensive review on its possible cosmological relevance and on the experimental efforts for detecting it as a relic particle, see for instance Ref. \cite{axion}. Among the particle candidates for CDM the most favorite one is certainly the Lightest Supersymmetric Particle (LSP), under the condition that it is weakly interacting. This candidate is discussed in the following section. \section{Supersymmetric dark matter particles} In order to be a dark matter candidate a particle has to be weakly interacting and stable (or, at least, long lived on cosmological time--scales). A very interesting perspective for such a candidate is offered, in the framework of supersymmetric theories with R-parity conservation, by the lightest susy particle (provided this is indeed weakly interacting). Different candidates have been considered: the neutralino \cite{neutralino} or the sneutrino\cite{snu} in gravity mediated models, the gravitino\cite{gravitino} or messenger fields\cite{messenger} in gauge mediated theories, the axino\cite{axino}, stable non--topological solitons (Q--balls)\cite{qballs}, heavy non--thermal relics\cite{zillas} or others. Among the different candidates, the most promising one turns out to be the neutralino, since its relic abundance may be sizeable, at the level required to explain the CDM content of the Universe and, at the same time, its detection rates may be accessible to experimental searches of different kinds. The phenomenology of neutralino dark matter has been studied extensively in the Minimal Supersymmetric extension of the Standard Model (MSSM) \cite{susy}. This model incorporates the same gauge group as the Standard Model and the supersymmetric extension of its particle content. The Higgs sector is slightly modified as compared to that of the Standard Model: the MSSM requires two Higgs doublets $H_1$ and $H_2$ in order to give mass both to down-- and up--type quarks and to cancel anomalies. After electroweak symmetry breaking, the physical Higgs fields consist of two charged particles and three neutral ones: two scalar fields ($h$ and $H$) and one pseudoscalar ($A$). The Higgs sector is specified at the tree level by two independent parameters: the mass of one of the physical Higgs fields and the ratio of the two vacuum expectation values, usually defined as $\tan\beta =\, <H_2> / <H_1>$. The supersymmetric sector of the model introduces some other free parameters: the mass parameters $M_1$, $M_2$ and $M_3$ for the supersymmetric partners of gauge fields (gauginos), the Higgs--mixing parameter $\mu$ and, in general, all the masses of the scalar partners of the fermions (sfermions) and all the trilinear couplings which enter in the superpotential. In the MSSM it is generally assumed that the gaugino masses are related by expressions induced by grand--unification. Specifically, the two parameters which are relevant for neutralino phenomenology are linked by: $M_1= (5/3) \tan^2 \theta_W M_2$. The other usual assumptions are that all slepton mass parameters are taken as degenerate to a common value $m_0$ and that all the trilinear couplings are vanishing except the ones of the third family which are set to a common value $A$. In summary, the free parameters of the model are six: $M_2$, $\mu$, $\tan\beta$, $m_A$, $m_0$ and $A$. The neutralinos are four mass--eigenstates defined as linear superpositions of the two neutral gauginos ($\tilde \gamma$ and $\tilde Z$) and the two neutral higgsinos ($\tilde H_1$ and $\tilde H_2$) \begin{equation} \chi = a_1 \tilde \gamma + a_2 \tilde Z + a_3 \tilde H_1 + a_4 \tilde H_2\;. \end{equation} \noindent The lowest--mass eigenstate may play the role of the lightest supersymmetric particle in the MSSM, and may then constitute the dark matter candidate in this model. It will be called the neutralino tout--court and its mass denoted by $m_\chi$. To classify the nature of the neutralino it is useful to define a parameter $P \equiv a_1^2 + a_2^2$; the neutralino is called a {\it gaugino}, when $P > 0.9$, a {\it higgsino} when $P < 0.1$ and {\it mixed} when $0.1 \leq P \leq 0.9$. The low--energy MSSM scheme is a phenomenological approach, whose basic idea is to impose as few model--dependent restrictions as possible. At a more fundamental level, it is natural to implement this phenomenological scheme within the supergravity (SUGRA) framework \cite{sugra}. One attractive feature of the ensuing model is the connection between soft supersymmetry breaking and electroweak symmetry breaking, which would then be induced radiatively. Usually, the low--energy phenomenology of SUGRA theories constitutes a subset of the susy configurations which are considered in the MSSM. A typical characteristic of SUGRA models is in fact the presence of relatively strong correlations among the low--energy parameters, correlation which is absent in the MSSM. In this report we will discuss neutralino dark matter phenomenology in the less constrained MSSM model. Results for SUGRA schemes can be found in the papers listed in Ref.\cite{sugra_dm} \subsection{Neutralino relic abundance} Neutralinos decouple from the hot plasma in the early Universe when they are no longer relativistic. Their present abundance is calculated by solving the Boltzmann equation for the evolution of the density of particle species\cite{kt} and can be written as: \begin{equation} \Omega_\chi h^2 = {\cal C} \, \frac{g_*^{1/2}(T_f)}{{g_*}_S (T_f)} \, \frac{1}{\langle \sigma_{\mbox{\rm\scriptsize ann}} v_r \rangle_{\mbox{\rm\scriptsize int}}} \, \end{equation} where \begin{equation} {\cal C} = \frac{s_0}{0.264\, \rho_c\, M_P} = 8.7 \cdot 10^{-11} \, \mbox{\rm GeV}^{-2} \, . \end{equation} In the previous Eqs., $g_*(T_f)$ and ${g_*}_S(T_f)$ denote the effective number of degrees of freedom for the energy density and for the entropy density, respectively, evaluated at the freeze--out temperature $T_f$; $\langle \sigma_{\mbox{\rm\scriptsize ann}} v_r \rangle_{\mbox{\rm\scriptsize int}}$ is the neutralino pair annihilation times the pair relative velocity, averaged over the neutralino thermal density distribution, integrated from the freeze--out temperature down to the present temperature; $s_0$ denotes the present entropy density, $\rho_c$ is the critical density and $M_P$ is the Planck mass. The critical quantity to be evaluated is the neutralino annihilation cross section, which, depending on the neutralino mass, can get contributions from the following final states : i) fermion--antifermion pair, ii) pair of neutral Higgs bosons, iii) pair of charged Higgs bosons, iv) one Higgs boson-one gauge boson, v) pair of gauge bosons. The diagrams contributing to the final state i) are Higgs--exchange Z--exchange diagrams in the s--channel, sfermion--exchange diagrams in the t--channel. For the other final states, the contributions come from Higgs and Z--diagrams in the s--channel, and either neutralinos or charginos exchange in the t--channel, depending on the electric charges of the final particles \cite{noi_omega,omegah2}. When the mass of the neutralino is close to the mass of another supersymmetric particle, the process of co--annihilation \cite{coann,suede_coann} can be present. In this case, the calculation of the annihilation cross section and of its thermal average has to take into account a large number competing interactions among the neutralino and its close--mass particles. In some special instances the relic abundance may be affected by co--annihilation effects by a sizeable amount\cite{suede_coann}. In Fig. 1 we show $\Omega_\chi h^2$ as a function of the neutralino mass $m_\chi$ \cite{noi_omega}. We present here, as well as in the following, all the results in terms of scatter plots which are obtained by varying the supersymmetric parameters inside wide ranges. Naturally, the parameter space is constrained by experimental limits on Higgs bosons and supersymmetric particles searches (for recent updates, see \cite{LEP}) and by measurements of rare processes, whose theoretical predictions are quite sensitive to supersymmetric contributions. At present, the most important is the decay $b \rightarrow s + \gamma$\cite{bsg_exp,bsg_theo}. The scatter plot of Fig. 1 is obtained by varying the supersymmetric parameters in the following ranges: $20\;\mbox{GeV} \leq M_2 \leq 500\;\mbox{GeV},\; 20\;\mbox{GeV} \leq |\mu| \leq 500\;\mbox{GeV},\; 80\;\mbox{GeV} \leq m_A \leq 1000\;\mbox{GeV},\; 100\;\mbox{GeV} \leq m_0 \leq 1000\;\mbox{GeV},\; -3 \leq {\rm A} \leq +3,\; 1 \leq \tan \beta \leq 50$. The figure shows only those configurations which provide a value for the relic abundance lower than a cosmological upper bound which has been conservatively set as $\Omega_\chi h^2 \leq 0.7$. The susy configurations which entail larger values of $\Omega_\chi h^2$ are excluded by the lower limit on the age of the Universe \cite{cdkk}. This constraint is rather restrictive on the susy parameter space. \subsection{Detection of relic neutralinos} Relic neutralinos would act as CDM during the process of galaxy formation. It is therefore conceivable that they may constitute all or part of the dark matter halo. These neutralinos would be clustered in the Galaxy and hence possess a matter density distribution and a velocity distribution which depend on the dynamics of the Galaxy formation and evolution\cite{BT}. Many different models have been discussed in the literature for the {\em matter density distribution} $\rho(\vec r)$ (see for instance Refs. \cite{nfw1,nfw2,kkbp}). This field is in rapid expansion, due to the high resolution simulations now at hand to investigate the structure of single galaxies \cite{moore}. In particular, these studies are expected to pin down the nature of the cuspy behaviour which appears to occur near the galactic center. Another very interesting possibility which is currently being investigated is that the halo could present a clumpy distribution of dark matter together with or in alternative to a smooth distribution. The uncertainties in the shape profile, combined with experimental uncertainties and the possibility that some fraction of the dark halo is made of baryonic dark matter in the form of MACHOS, implies that the local value of the matter density $\rho_l = \rho(\vec r_\odot)$ is somewhat uncertain. At present, its most conservative range of variability can be set as \cite{turner_rhol}: 0.1 GeV cm$^{-3}$ $\leq \rho_l \leq 0.7$ GeV cm$^{-3}$. The quantity $\rho_l$ refers to the total dark matter density of the galactic halo. the neutralino local density is, in general, a fraction of it, i.e. $\rho_\chi = \xi \rho_l$, with $\xi \leq 1$. No exact way to evaluate the quantity $\xi$ is currently available. A usual procedure is to consider the galactic halo as composed entirely of neutralinos if, on the average in the Universe, neutralinos are abundant enough to explain all the CDM which is observed at least on the galactic scale. This happens when $\Omega_\chi h^2$ is larger than a value $(\Omega h^2)_{\rm min}$ derived from astrophysical observations. In this case it is possible to set $\xi = 1$. If, on the contrary, $\Omega_\chi h^2 < (\Omega h^2)_{\rm min}$, neutralinos are not able to explain all the CDM, even the one which is needed at the galactic scale, and therefore also locally in our Galaxy we expect them to give only a fractional contribution $\xi$ to $\rho_l$. In this case, the simplest choice is to set: $\xi = \Omega_\chi h^2 / (\Omega h^2)_{\rm min} $. The quantity $ (\Omega h^2)_{\rm min}$ is estimated to lie in the range $0.01 \lsim (\Omega h^2)_{\rm min} \leq 0.2$. The {\em velocity distribution} $f(\vec v)$ of dark matter is usually assumed to be a Maxwellian distribution \cite{BT} (as seen from the galactic rest frame), with a velocity dispersion $v_{\rm rms}$ which is directly related to the asymptotic flat rotational velocity $v_\infty$ as: $v_{\rm rms} = \sqrt{3/2}v_\infty$. In our Galaxy, the rotational velocity appears to be already flat at the local position, and therefore we set $v_{\infty} = v_{\odot}$. The experimental determination of the local rotational velocity is: $v_{\odot} = 230 \pm 50$ Km s$^{-1}$ (90 \% C.L.)\cite{koch}. The distribution is also truncated by an escape velocity $v_{\rm esc}$ which falls in the range: $v_{\rm esc} = 450 \div 650$ Km s$^{-1}$ (90 \% C.L.)\cite{leonard,cud}. Modifications to the standard Maxwell--Boltzmann form have been examined \cite{BT,evans}, but the problem of determining the correct form of the distribution of the velocities in the halo has no clear and simple solution at present, both theoretically and observationally. Also the possibility that the halo could possess a bulk rotation has been considered in the literature\cite{fv_rot}. Due to the possibility that neutralinos are present in the halo of our Galaxy, it is of great interest, both for astrophysics and particle physics, to search for techniques capable of detecting these dark halo particles. Several methods have been proposed and in the following we will briefly review the ones which, at present, appear to be more promising. \subsection{Direct detection of relic neutralinos} The most direct possibility to detect dark matter particles is to look for their scattering with the nuclei of a low--background detector\cite{witten}. The interaction of slow halo neutralinos of mass $m_\chi \gsim 25$ GeV with a detector produces the recoil of a nucleus with energy $E_R$ of the order of few to tens keV. The recoil energy can be measured by means of various experimental techniques with different nuclear species, like Ge, NaI, Xe, CaF$_2$, TeO$_2$\cite{taup_direct}. The relevant quantity to be calculated and compared with the experimental measurements is the differential detection rate \begin{equation} S_0(E_R) \equiv \frac {dR}{dE_R}=N_{T}\frac{\rho_{\chi}}{m_{\chi}} \int \,d \vec{v}\,f(\vec v)\,v \frac{d\sigma}{dE_{R}}(v,E_{R}) \label{rate} \label{eq:diffrate0} \end{equation} where $N_T$ is the number of the target nuclei per unit of mass, $\rho_\chi$ is the local neutralino matter density, $\vec v$ and $f(\vec v)$ denote the neutralino velocity and velocity distribution function in the Earth frame ($v = |\vec v|$). The nuclear recoil energy is given by $E_R={{m_{\rm red}^2}}v^2(1-\cos \theta^*)/{m_N}$, where $\theta^*$ is the scattering angle in the neutralino--nucleus center--of--mass frame, $m_N$ is the nuclear mass and $m_{\rm red}$ is the neutralino--nucleus reduced mass. Finally, $d\sigma/dE_R$ is the neutralino--nucleus differential cross section, which has a coherent contribution due to Higgs-- and squark--exchange, and a spin dependent one which originates from the exchange of the Z boson and squarks. The coherent cross section is usually largely dominant over the spin--dependent one. Eq.(\ref{eq:diffrate0}) refers to the situation of a monoatomic detector, like the Ge detectors. For more general situations, like for instance the case of NaI, the generalization is straightforward. {}From those experimental data on the nuclear recoil spectrum which do not provide a signal--to--background discrimination, upper limits to the neutralino--nucleus cross section as a function of the neutralino mass may be set by employing Eq.(\ref{eq:diffrate0}) \cite{bdmsbi}. In the case of coherent interaction, Fig. 2 shows, as a solid line, the present most stringent upper limit\cite{DAMA_uplim} on the quantity $\xi \sigma^{(\rm nucleon)}_{\rm scalar}$ vs. $m_\chi$, where $\sigma^{(\rm nucleon)}_{\rm scalar}$ denotes the scalar elastic cross section of a neutralino off a nucleon. The astrophysical parameters are chosen as: $\rho_l = 0.3$ GeV cm$^{-3}$, $v_0 = 220$ Km s$^{-1}$, $v_{\rm esc} = 550$ Km s$^{-1}$ and $ (\Omega h^2)_{\rm min} = 0.01$. In this plot, we also show a scatter plot of the same quantity calculated in the MSSM\cite{noi_diretta,diretta_theo}. The susy configurations have been varied in the ranges quoted in Sect. 6.1. We stress that the cosmological bound $\Omega_\chi h^2 \leq 0.7$ has been applied \cite{note_damour}. In the case of direct detection, a typical signature consists in the annual modulation of the detection rate\cite{ann_mod_th}. During the orbital motion of the Earth around the Sun, the change of direction of the relic particle velocities with respect to the detector induces a time dependence in the differential detection rate, i.e. $S(E_R,t) = S_0 (E_R) + S_m (E_R) \cos [\omega (t-t_0)]$, where $\omega = 2\pi/365$ days and $t_0 = 153$ days (June 2$^{\rm nd}$). $S_0 (E_r)$ is the average (unmodulated) differential rate defined in Eq.(\ref{eq:diffrate0})and $S_m (E_R)$ is the modulation amplitude of the rate. The relative importance of $S_m (E_R)$ with respect to $S_0 (E_R)$ for a given detector, depends both on the mass of the dark matter particle and on the value of the recoil energy where the effect is looked at. Typical values of $S_m (E_R)/S_0 (E_R)$ for a NaI detector range from a few percent up to $\sim$ 15\%, for neutralino masses of the order of 20--80 GeV and recoil energies below 8--10 KeV. The search for the annual modulation effect is currently undertaken by the DAMA/ NaI Collaboration\cite{DAMA_longrep}. The analysis of their data over two years of data--taking provides the indication of a modulated signal \cite{DAMA}. This result can be shown as the closed contour in the $\xi \sigma^{(\rm nucleon)}_{\rm scalar}$ vs. $m_\chi$ plane displayed in Fig. 2. The region inside the contour is compatible with a modulation signal at 2--$\sigma$ level. The contour takes into account also the uncertainties in astrophysical velocities, as discussed in \cite{bellietal}. Fig. 2 shows that there exist susy configurations which would be able to explain such an effect. In the papers of Refs.\cite{bellietal,noi_DAMA} it has been proved that some of these configurations are explorable at accelerators and/or by WIMP indirect experiments and that the relevant relic neutralinos might behave as major components of cold dark matter. For an analysis of these configurations in a SUGRA scheme, see also Ref.\cite{an}. \subsection{Indirect detection: neutrino fluxes from Earth and Sun} Other ways of detecting dark matter neutralinos rely on the possibility to detect the products of neutralino annihilations. One perspective is to observe a neutrino signal coming from celestial bodies, namely Earth and Sun, where the neutralinos may have been captured and accumulated during the lifetime of the macroscopic body. The sequence of the physical processes which could produce these signals are: i) capture of relic neutralinos by the macroscopic bodies; ii) subsequent accumulation of these particles in a region around the centre of these celestial objects; iii) pair annihilation of the accumulated neutralinos which would generate, by decay of the particles produced in the various annihilation final states, an output of high--energy neutrinos. Since the process of annihilation takes place inside a medium (i.e., the interior of the Earth or the Sun), the annihilation process is perturbed as compared to the annihilation in vacuum. This effect can be effectively taken into account\cite{ritz-seckel} by neglecting the contributions of the light quarks directly produced in the annihilation process or in the hadronization of heavy quarks and gluons, because these light particles stop inside the medium (Sun or Earth) before their decay. For the case of the Sun, also the energy loss of the heavy hadrons in the solar medium and the energy loss of neutrino themselves have to be considered\cite{ritz-seckel}. The differential neutrino flux is then calculated as \begin{equation} \frac{dN_\nu}{dE_\nu} = \frac{\Gamma_A}{4\pi d^2} \sum_{F,f} B^{(F)}_{\chi f}\frac{dN_{f \nu}}{dE_\nu} \end{equation} where $\Gamma_A$ denotes the annihilation rate, $d$ is the distance of the detector from the source (i.e. the center of the Earth or the Sun), $F$ denotes the neutralino pair annihilation final states, $B^{(F)}_{\chi f}$ denotes the branching ratios into heavy quarks, $\tau$ lepton and gluons in the channel $F$. The spectra $dN_{f \nu}/dE_{\nu}$ are the differential distributions of the neutrinos generated by the $\tau$ and by hadronization of quarks and gluons and the subsequent semileptonic decays of the produced hadrons. A detailed calculation of these spectra is usually performed by means of a Montecarlo simulation\cite{noi_nuflux}. The spectra due to heavier final states, i.e. Higgs bosons, gauge bosons and t quark, can be computed analytically by following their decay chain down to the production of a b quark, c quark or a tau lepton, where the result of the Montecarlo simulation can be applied\cite{noi_nuflux,altri_nuflux}. The neutrino flux may be detected in a neutrino telescope by measuring the muons which are produced by $\nu_\mu$ and ${\bar \nu}_\mu$ interactions in the rock around the detector and then traverse it upwardly. Therefore, the signal consists of a flux of up--going muons, which is computed as \begin{equation} \frac{d N_\mu}{d E_\mu} = N_A \int^\infty_{E_\mu^{\rm th}} d E_\nu \int_0^\infty dX \int_{E_\mu}^{E_\nu} d {E'_\mu }\,\, P_{\rm surv}(E_\mu,E'_\mu; X)\,\, \frac{d \sigma_\nu (E_\nu,E'_\mu)}{d E'_\mu} \,\, \frac{d N_\nu}{d E_\nu}\, , \label{upmu_flux} \end{equation} where $X$ is the muon range in the rock, $d \sigma_\nu (E_\nu,E'_\mu) / d E'_\mu$ is the charged current cross--section for the production of a muon of energy $ E'_\mu$ from a neutrino of energy $E_\nu$ and $P_{\rm surv}(E_\mu,E'_\mu; X)$ is the survival probability that a muon of initial energy $E'_\mu$ will have a final energy $E_\mu$ after propagating along a path--length $X$ inside the rock which surrounds the detector. The function $P_{\rm surv}(E_\mu,E'_\mu; X)$ therefore takes into account the energy losses of muons in the rock. In Eq.(\ref{upmu_flux}), $E_\mu^{\rm th}$ is the detector threshold energy, which for current neutrino telescopes like MACRO and Baksan is of about 1-2 GeV, and this is quite suitable for neutralino indirect detection, especially for neutralinos lighter than about 100 GeV. Large--area neutrino telescopes with higher threshold energies (above a few tens of GeV) are more suitable for heavier relic particles. Experimentally, one searches for a statistically significant excess of up--going muons above the muon flux originated from atmospheric neutrinos. The different angular behaviour of the signal with respect to the atmospheric neutrino background, which has a rather flat distribution as a function of the zenith angle, is the handle for the signal--to--background discrimination. Clearly, the flux from the Sun can be pointed at directly towards the direction of the Sun. In the case of the flux from the Earth, the process of accumulation of neutralinos induces a rather peaked distribution of the neutrino source around the Earth's center. Indeed, the angular distribution is \begin{equation} G(\theta) \simeq 4 m_\chi \alpha \exp (-2 m_\chi \alpha \sin^2 \theta) \end{equation} where $\theta$ is the zenith angle and $\alpha = 1.76$ GeV$^{-1}$. This means that for neutralinos (which are heavier than $\sim$ 25 GeV) the signal is produced inside a region whose angular extension is less than about 10 degrees. The experimental searches at neutrino telescopes have found no muon excess so far and therefore upper limits on the muon flux have been set. The solid line in Fig. 3a is the current most stringent experimental upper limit from the MACRO Collaboration\cite{MACRO,taup_direct} for the neutrino flux from the center of the Earth. Fig. 3b refers to the flux from the Sun. Again, superimposed to the experimental limits, we show in Fig. 3a (Earth) and Fig. 3b (Sun) the susy scatter plot for the up--going muon signal $\Phi_\mu^{\rm Earth}$ and $\Phi_\mu^{\rm Sun}$. The susy configurations have been varied in the ranges quoted in Sect. 6.1. The astrophysical parameters are $\rho_l = 0.3$ GeV cm$^{-3}$, $v_0 = 220$ Km s$^{-1}$, $v_{\rm esc} = 550$ Km s$^{-1}$ and the cosmological bound $\Omega_\chi h^2 \leq 0.7$ has been applied. \subsection{Indirect detection: antimatter and gamma rays} The annihilation process of dark matter neutralinos may take place also directly in the galactic halo. In this case, many different signals other than neutrinos are possible. These signals, at variance with the signals previously discussed, which depend only on local galactic properties, depend directly on the matter distribution of neutralinos over the whole Galaxy. Moreover, the propagation inside the Galaxy of the particles which constitute the signal is perturbed by the Galaxy itself, like, for instance, interactions with the interstellar medium or, in the case of charged particles, diffusion in random magnetic fields. One of the most interesting possibility is the production of antimatter from neutralino annihilation in the halo. The fluxes have been calculated for production of antiprotons, antideuteron and positrons. {\em Antiprotons}\cite{noi_pbar,others_pbar} and {\em Antideuterons}\cite{dbar} can be produced by the decay and hadronization of the final state particles of the annihilation process. The differential rate per unit volume and unit time for the production of $ \bar p$'s from neutralino pair annihilation is defined as \begin{equation} q_{\bar p}^{\rm susy}(T_{\bar p}) = \, <\sigma_{\rm ann} v> \left( \frac{ \rho_\chi (\vec r)}{m_\chi} \right)^2\, \sum_{F,h} B^{(F)}_{\chi h} \frac{dN^{h}_{\bar p}}{dT_{\bar p}} \, , \label{eq:source} \end{equation} where $<\sigma_{\rm ann} v>$ denotes the average over the galactic velocity distribution function of neutralino pair annihilation cross section $\sigma_{\rm ann}$ multiplied by the relative velocity $v$ of the annihilating particles, $T_{\bar p}$ is the antiproton kinetic energy and $B^{(F)}_{\chi h}$ is the branching ratio for the production of quarks or gluons $h$ due to the decay of the particles produced by neutralino annihilation into the final state $F$. Finally, $dN^h_{\bar p}/dT_{\bar p}$ is the differential energy distribution of the antiprotons generated by hadronization of quarks and gluons. Notice that the rate depends on the square of the mass distribution function of neutralinos in the galactic halo $\rho_\chi (\vec r)$. The rate of production of $\bar D$ is clearly analogous to Eq.(\ref{eq:source}). After being produced, antimatter propagates inside the Galaxy and it experiences both diffusion in the galactic magnetic field and energy losses, due to ionization, scattering, collision and others. The propagation of antiprotons in the galactic medium is properly calculated in a diffusion model where the Galaxy is described as composed of two zones, one for the disk and the other for the halo. The diffusion equation which governs the behaviour of antiprotons is \begin{equation} \vec{\nabla} \cdot \left( K \; \vec{\nabla} \, \mbox{$\psi_{\bar{\rm p}}$}\right) \; - \; 2 h \delta(z) \, \Gamma_{\bar{\rm p}} \, \mbox{$\psi_{\bar{\rm p}}$}\; + \; 2 h \delta(z) \, \mbox{$q_{\bar{\rm p}}^{\rm susy}$}(r) \; - \; 2 h \delta(z) \, \frac{\partial}{\partial E} \left\{ b(E) \mbox{$\psi_{\bar{\rm p}}$}\right\} \; = \; 0 \;\; , \label{DIFFUSION_PBAR_DISK} \end{equation} where $\psi_{\bar p}$ denotes the antiproton density, $K$ is the diffusion coefficient, $h$ is the height of the galactic disk and $\Gamma_{\bar{\rm p}}$ is the collision probability with the interstellar medium and $b(E)$ describes energy losses. An analogous equation holds for the antideuterons. Solution of the diffusion equations gives the antiproton (or antideuterons) flux at the heliosphere boundaries (interstellar flux) as \begin{equation} \Phi_{\bar{\rm p}} ( \odot , T_{\bar{\rm p}} ) \,\, = \,\, <\sigma_{\rm ann} v> \, \frac{v_{\bar p}}{4 \pi} \, \left ( \frac{ \rho_l}{m_\chi} \right )^2 \, \psi_{\bar{\rm p}}^{\rm eff} ( \odot , T_{\bar{\rm p}} ) \cdot \sum_{F,h} B^{(F)}_{\chi h} \frac{dN^{h}_{\bar p}}{dT_{\bar p}}\, , \label{SUSY_ANALYSIS} \end{equation} where $\psi_{\bar{\rm p}}^{\rm eff} ( \odot , T_{\bar{\rm p}} ) $ is obtained by solving Eq.(\ref{DIFFUSION_PBAR_DISK}). Antimatter subsequently enters the heliosphere where it propagates against the solar wind before arriving at the Earth. The effect induced by the solar wind (solar modulation) is quite important at low kinetic energies, and introduces a time dependence into the calculation, since it is correlated with the 11 year solar cycle. Both antiprotons and antideuterons can be produced also by the interaction of primary cosmic rays on the interstellar medium. This secondary antimatter fluxes constitute a background for the susy signal. The background fluxes have a different energy behaviour as compared to the the ones of susy origin. In particular, the low energy tail of the energy spectrum is the most interesting place to look at, since the signals have a somewhat flat behaviour, while the secondary fluxes are depressed by kinematical reasons \cite{secondary}. This effect is stronger for antideuterons than for antiprotons. Therefore, the antideuteron flux presents some advantages of discrimination over background with respect to the antiproton flux, even if its smallness makes the detection harder to be achieved. Fig. 4 show the antiproton flux\cite{noi_pbar} $\Phi_{\bar p}$ vs. the neutralino mass for the susy configurations introduced in Sect. 6.1. The flux is calculated for a $\bar p$ kinetic energy $T_{\bar p} = 0.24$ GeV, to conform to the first BESS energy bin, and for a solar modulation phase close to the BESS data--taking periods. The horizontal line is the present upper limit derived from the BESS 95 and BESS 97 data\cite{BESS97}. We notice that antiproton measurement are already able to exclude some susy configurations which would imply a too large $\bar p$ flux at low kinetic energies. Fig. 5 shows the antideuteron flux\cite{dbar} $\Phi_{\bar D}$ vs. the neutralino mass for the same susy models. The flux is calculated for a $\bar D$ kinetic energy $T_{\bar D} = 0.24$ GeV/nucleon, at the maximum of solar activity. The horizontal line is the sensitivity level which could be achieved by the AMS detector during the flight on space station. {\em Positrons}\cite{suede_positron} are produced again from the decay chain of the neutralino annihilation products. It is also possible to produce directly a pair of monocromatic electron--positron pair. The branching ratio for this process is usually very small, but some susy models can present a production rate strong enough to be at the level of the detector sensitivity\cite{suede_positron}. The calculation of the positron flux is analogous to the one for antiprotons or antideuterons, with the inclusion of additional energy loss mechanisms: inverse Compton scattering on starlight and cosmic microwave background, and synchrotron radiation emission in the galactic magnetic field. The IS positron flux is then affected by the solar wind before coming to the Earth where it can be detected. Fig. 6, taken from Ref.\cite{suede_positron}, shows a scatter plot of the positron flux $\Phi_{e+}$ integrated in the energy range 8.9 -- 14.8 GeV, to conform to one of the HEAT data bins. The susy parameters have been varied in the ranges: $0\;\mbox{GeV} \leq |M_2| \leq 5000\;\mbox{GeV},\; 0\;\mbox{GeV} \leq |\mu| \leq 5000\;\mbox{GeV},\; 0\;\mbox{GeV} \leq m_A \leq 10000\;\mbox{GeV},\; 100\;\mbox{GeV} \leq m_0 \leq 30000\;\mbox{GeV},\; -3 \leq {\rm A} \leq +3,\; 1 \leq \tan \beta \leq 60$. Only configuration with $0.025 \leq \Omega_\chi h^2 \leq 1$ are shown on the plot. The horizontal dashed line denotes the value of the background positrons of secondary origin calculated in Ref.\cite{moskalenko}. The horizontal band represents the HEAT 94 data\cite{HEAT}. The last possibility we consider here is the production of {\em diffuse gamma rays} or a {\em gamma line} (see Refs.\cite{gamma,gamma_bbm,line_bba,bub,suede_clump,gamma_annihil} and references quoted therein). Diffuse photons are produced mainly through the decay of neutral pions originated from the hadronization of the neutralino annihilation products. A monochromatic gamma line, instead, is produced through the loop--processes $\chi \chi \longrightarrow \gamma \gamma$ and $\chi \chi \longrightarrow Z \gamma$. In this case, the gamma line would constitute a particularly nice signature, since it is practically background free. In both cases, the fluxes are usually rather low and, in order to have fluxes at the level of the detector sensitivities, some matter over-density is needed, like for example a singular dark matter halo\cite{gamma_bbm,bub} or a clumpy matter distribution\cite{suede_clump}. Fig. 7, taken from Ref.\cite{suede_clump}, shows a scatter plot of the gamma ray flux $\Phi_{cont, \gamma}$ integrated above $E_\gamma = 1$ GeV. The susy parameters have been varied in the same ranges defined above for the positron flux. Only configurations with $0.025 \leq \Omega_\chi h^2 \leq 0.5$ are shown on the plot. The horizontal line denotes the integrated gamma ray flux measured by EGRET\cite{EGRET}. Fig. 8 from Ref. \cite{bub} shows the perspectives of a number of Air Cherenkov Telescopes for the measurements of the expected $\gamma$--ray line. \section{Conclusions} As we have seen above, the question of dark matter in the Universe presents a large number of intriguing facets of relevance for cosmology, astrophysics and particle physics. It represents a field in a very fast expansion, because of an impressive development in experimental activities as well as in theoretical investigations. Let us mention just some of the most promising avenues: a) Further observations and analyses of high--$z$ Supernovae, of the Cosmic Microwave Background Radiation, and of the time evolution of the number density of clusters are expected to provide more conclusive information on $\Omega_m$ and $\Omega_{\Lambda}$. b) New numerical simulations of cosmological structures should give a unique information about the (hot/cold) composition of dark matter and about crucial details on the dark matter density profile in single galaxies. c) Further accumulation of data in WIMP direct detection are expected to play a fundamental role in the process of the identification of dark matter particle candidates. Present experimental results have already allowed the pinning down of a sector of the supersymmetric parameter space, part of which can be explored at accelerator and by WIMP indirect searches. It has been proved that the relevant relic neutralinos might behave as major components of cold dark matter. No doubt that the connection of particle physics with the dark matter problem in the Universe is one of the most exciting and far--reaching field in astroparticle physics. \bigskip \bigskip \begin{center} {\bf Acknowledgments} \end{center} This work was supported by DGICYT under grant number PB95--1077 and by the TMR network grant ERBFMRXCT960090 of the European Union. \bigskip \bigskip
\section{Introduction} Any laboratory on the Earth simultaneously experiences both acceleration and gravity. Einstein's equivalence principle states that, neglecting curvature effects, the acceleration and gravitation are indistinguishable. The success of General Relativity is a testament to the accuracy of this statement in the classical regime. The equivalence principle has also been tested in quantum mechanical phenomena. First, Colella et al.\@~\cite{ref-colella} performed their celebrated neutron interferometry experiment, observing a gravitationally-induced phase shift. Using similar apparatus, Bonse and Wroblewski~\cite{ref-bonse} found the same phase shift when their interferometer was uniformly accelerated. Together, these experiments provided some confirmation of the equivalence principle in quantum mechanics. These two neutron interferometry experiments did not measure any spin-dependent effects, leaving open the question of whether spin may be affected differently by acceleration and uniform gravitation. Varj\'{u} and Ryder~\cite{ref-varju} have suggested a particular spin effect that may be able to distinguish between the two. They compared two Dirac (spin 1/2) Hamiltonians. One was the Hamiltonian for a particle in a Schwartzchild gravitational field, drawing on earlier calculations by Fischbach, et al.\@~\cite{ref-fischbach}. The other Hamiltonian represented a Dirac particle subjected to a uniform acceleration, as found by M.~S.~Altschul~\cite{ref-maltschul} and later by Hehl and Ni~\cite{ref-hehl}. An apparent difference between these Hamiltonians calls into question whether the equivalence principle holds for spin effects. We will show that that an additional accelerational effect could eliminate the difference between the two Hamiltonians, restoring the equivalence principle. However, it is natural that this effect has not yet been observed, since it acts on very long length scales. \section{Comparison of Hamiltonians} In comparing uniform gravitation and acceleration, we will evaluate the Hamiltonians to first order. For the acceleration, that means to first order in ${\bf a}$, the acceleration vector. For the gravitational Hamiltonian, we will work to first order in ${\bf g}$, where $g_{i}=-\frac{\partial \Phi}{\partial x^{i}}c^{2}$. Here, $\Phi=\frac{GM}{c^{2}r}$, so ${\bf g}=-\frac{\Phi c^{2}} {r^{2}}{\bf x}$ is just the normal gravity vector. Each of these Hamiltonians may be found by using three successive Foldy-Wouthuysen transformations~\cite{ref-foldy}. The gravitational Hamiltonian found by Varj\'{u} and Ryder takes the form \begin{eqnarray} H&=&\beta mc^{2}-\beta m{\bf g}\cdot {\bf x}+\frac{\beta}{2m}p^{2}-\frac{\beta} {2mc^{2}}{\bf p}\cdot({\bf g}\cdot {\bf x}){\bf p}\nonumber \\ & &+\frac{\hbar\beta}{2mc^{2}}\mbox{{\boldmath $\sigma$}}\cdot({\bf g}\times {\bf p})-\frac{\beta}{mc^{2}}({\bf p}\cdot {\bf g})({\bf x}\cdot {\bf p}). \label{eq-ghamil} \end{eqnarray} The accelerational Hamiltonian has been found in two different (and both flawed) ways. The result was \begin{eqnarray} H&=&\beta mc^{2}+\beta m{\bf a}\cdot {\bf x}+\frac{\beta}{2m}p^{2}+\frac{\beta} {2mc^{2}}{\bf p}\cdot({\bf a}\cdot {\bf x}){\bf p}\nonumber \\ & &+\frac{\hbar\beta}{4mc^{2}}\mbox{{\boldmath $\sigma$}}\cdot({\bf a}\times {\bf p}). \label{eq-ahamil} \end{eqnarray} The equivalence principle indicates that these two Hamiltonians should be identical when we set ${\bf a}=-{\bf g}$, except for tidal terms in (\ref{eq-ghamil}). On initial inspection, this does not appear to be the case. There are two differences. The last term of (\ref{eq-ghamil}) is entirely absent from (\ref{eq-ahamil}). Although only first order in ${\bf g}$, this term is actually a tidal term, as Varj\'{u} and Ryder demonstrated, so its absence from (\ref{eq-ahamil}) is expected. The other difference between the Hamiltonians less easily dismissed. While (\ref{eq-ghamil}) includes the term \begin{equation} \label{eq-gterm} \frac{\beta\hbar}{2mc^{2}} \mbox{{\boldmath $\sigma$}}\cdot({\bf g}\times {\bf p}), \end{equation} (\ref{eq-ahamil}) contains \begin{equation} \label{eq-aterm} \frac{\beta\hbar}{4mc^{2}} \mbox{{\boldmath $\sigma$}}\cdot({\bf a}\times {\bf p}). \end{equation} When we equate ${\bf a}=-{\bf g}$, these two terms differ by a factor of $-2$. It appears that this difference could allow an observer to distinguish between the acceleration and the gravitation. However, our closer analysis indicates that this may not be the case. \section{Anomalous Acceleration Moment} The Hamiltonian (\ref{eq-ahamil}) was first derived by purely electromagnetic means, for a charged particle in a uniform electric field. Then (\ref{eq-aterm}) enters as a charge-dependent term, vanishing for a neutral particle. The behavior of (\ref{eq-aterm}) in this regard is the same as that of the spin magnetic moment term in a uniform magnetic field ${\bf B}$. However, we know that the spin magnetic moment is actually an additional parameter of the Dirac Hamiltonian; it is not zero for neutral particles with internal structure. Since ${\bf a}$ and ${\bf B}$ have the same three-dimensional vector group structure, there should be an analogous parameter in the accelerational Hamiltonian. In fact, when we looks at the infinitesimal generator for an acceleration, we may identify this new parameter. As shown in~\cite{ref-maltschul}, the generator has the form \begin{equation} \label{eq-chisimple} \mbox{{\boldmath $\chi$}}=\mbox{{\boldmath $\chi$}}_{\Gamma^{0}}+ \frac{\beta\hbar}{4c^{2}}\,\mbox{\boldmath $\sigma$}\times\mbox {{\boldmath $\nabla$}}_{p}+f_{\chi}. \end{equation} For small values of ${\bf a}$ this generates the Hamiltonian (\ref{eq-ahamil}) from the free particle Hamiltonian $H_{0}$ by $H=H_{0}+[{\bf a}\cdot \mbox{{\boldmath $\chi$}},H_{0}]$. Under this prescription, the first term of {\boldmath $\chi$} generates the ``classical'' potential energy $\beta m {\bf a\cdot x}$, and $f_{\chi}$ commutes with $H_{0}$. The second term, containing {\boldmath $\sigma$}, is the one of interest. We will call this term $\mbox{{\boldmath $\chi$}}'$. It is $\mbox{{\boldmath $\chi$}}'$ that generates the problematic term (\ref{eq-aterm}). From the form of (\ref{eq-chisimple}), it is clear that there is a free parameter. The coefficient of {\boldmath $\chi'$} may be freely adjusted without affecting any lower-order terms in the Hamiltonian. An ``anomalous acceleration moment,'' $\mu_{a}$ may be introduced. The presence of $\mu_{a}$ changes the generator according to \begin{equation} \mbox{{\boldmath $\chi$}}'=(1+\mu_{a})\frac{\beta\hbar}{4c^{2}}\, \mbox{\boldmath $\sigma$}\times\mbox{{\boldmath $\nabla$}}_{p}. \end{equation} The variation of $\mu_{a}$ creates a one-parameter family of representations of the acceleration group. It remains to consider for what particles this anomalous contribution may be nonzero and significant. For charged particles, the contribution is certainly not significant. Attempts to accelerate such particles uniformly are prevented by the emission of electromagnetic radiation and the induced radiative reaction force~\cite{ref-jackson}. For neutral particles, this problem does not appear, so we shall consider neutrons. From the pure Dirac standpoint, the neutron should have no magnetic moment. However, there is a neutron magnetic moment caused by QCD. Similarly, there may be an nonzero anomalous acceleration moment for the neutron. In fact, it is expected to be nonzero, for the following reason. The value $\mu_{a}=0$ is required by the conditions of Lorentz invariance and charge conservation~\cite{ref-jaffe}. However, these conditions do not hold under acceleration. Since a charge density causes acceleration, the charge and acceleration operators do not commute. More specifically, the quark operators $\psi_{q}^{\dag}$ that compose a neutron do not commute with acceleration, so the quark content of an accelerated neutron is not well-defined. We may estimate the effective commutator from the fact that a free quark would induce a force $F\approx 10^{22}$~GeV/cm~\cite{ref-rolnick} on every differently-colored quark in the universe. This yields an order of magnitude of \begin{equation} |[a,\psi_{q}^{\dag}]| \sim\frac{2}{3}\frac{F}{m_{q}}\approx 10^{36}\, {\rm cm\cdot s^{-2}}, \end{equation} since the quark mass $m_{q}$ is a few MeV\,$\cdot\,c^{-2}$ for the $u$ and $d$ quarks. The lowest-order opportunity for terms stemming from this commutator to enter the Hamiltonian is through $\mu_{a}$. This noncommutation is what interferes with Hehl and Ni's derivation of the accelerational Hamiltonian, in which they simply replace the partial derivatives in the Dirac equation with covariant derivatives. Although Varj\'{u} and Ryder use this same substitution to find the gravitational Hamiltonian, the noncommutation of ${\bf g}$ with the quark operators lies in the realm of quantum gravity, and so should be negligible. So the existence of $\mu_{a}$ allows for the restoration of the equivalence principle. In fact, the equivalence principle provides a specific prediction of its value, $\mu_{a}=-3$. \section{Length Scale for the Acceleration Moment} It is important to investigate the dynamical effects caused by $\mu_{a}$. We shall begin this process by examining the length scale on which $\mbox{{\boldmath $\chi$}}'$ operates. The order of magnitude of $\mu_{a}$ will probably not deviate too much from unity, since it is a dimensionless parameter of nuclear structure. This is certainly true if $\mu_{a}$ takes the value indicated by the equivalence principle. To get a length scale from $\mbox{{\boldmath $\chi$}}'$, we must insert the quantum-mechanical prescription that $i\hbar\mbox{{\boldmath $\nabla$}}_{p}= {\bf x}$, to get \begin{equation} \label{eq-chiwithx} \mbox{{\boldmath $\chi$}}'=(1+\mu_{a})\frac{\beta}{4ic^{2}}\, \mbox{\boldmath $\sigma$}\times{\bf x}. \end{equation} The $i$ disappears when we take $\exp(i{\bf a}\cdot\mbox{{\boldmath $\chi$}}) \approx 1+i{\bf a}\cdot\mbox{{\boldmath $\chi$}}$. The terms multiplying ${\bf x}$ in $\mbox{{\boldmath $\chi$}}'$ give us a length scale for the action of this effect in the Hamiltonian. Although the generator is only correct for infinitesimal accelerations, it should give the right dependence for the length scale. We drop $\beta$ and {\boldmath $\sigma$}, since they only take the values $\pm 1$. This leaves \begin{displaymath} \frac{|1+\mu_{a}|}{4c^{2}}\,{\bf x}. \end{displaymath} Since the original dimensionless quantity for generating the Hamiltonian was ${\bf a}\cdot\mbox{{\boldmath $\chi$}}$, the length scale must depend on $a=|{\bf a}|$. The final expression for determining the length scale is then \begin{equation} {\bf a}\cdot\mbox{{\boldmath $\chi$}}'\sim\frac{|1+\mu_{a}|}{4c^{2}}\,{\bf |x||a|}. \end{equation} Therefore, the characteristic length scale for the accelerated neutron is \begin{equation} x_{a}=\frac{4c^{2}}{|1+\mu_{a}|a}. \end{equation} In cgs units, this has the numerical value of \begin{equation} \label{eq-scale} x_{a}=\frac{3.6\times 10^{21}\, {\rm cm}}{|1+\mu_{a}|a}. \end{equation} For reasonable values of $a$ and $|1+\mu_{a}|$, the scale of $x_{a}$is astrophysical. A natural place to look for this effect would be the solar wind, since the solar wind contains many neutrons and it undergoes continuous acceleration. The gravitational acceleration at the surface of the sun is $a_{\odot}=GM_{\odot}/R_{\odot}^{2}=2.7\times 10^{4}$~cm\,$\cdot\,$s$^{-2}$. If $|1+\mu_{a}|=2$, this indicates a length of $x_{a}=6.6\times 10^{16}$~cm or 0.036~LY. This is a substantial distance. Moreover, this is not the flight distance of the neutrons from the sun. Instead, it represents their retardation by the sun's gravitational attraction. The actual flight distance would be even larger, and the sun's gravitational acceleration is certainly not uniform over the distance. Thus, it is almost certainly not possible to observe any effect on the solar wind neutrons. Since the length scale due to the sun's gravitational acceleration is too large, it is natural to consider situations in which $a$ is substantially greater. In particular, we will look at the acceleration at the surface of a neutron star, with mass $1.5M_{\odot}$ and radius $10^{6}$ cm. For this body, $a_{N\- S}=2.0\times 10^{14}$~cm\,$\cdot$\,s$^{-2}$. In this case, the value of $x_{a}$ is $9.1\times 10^{6}$~cm, substantially smaller. However, this is still too large for the gravitational field to be nearly uniform. In any case, it is unlikely that particles at a neutron star's surface would be free from far stronger interactions, so the effect is almost certainly not significant in this case, either. These calculated length scales indicate that the term of the Hamiltonian containing $\mu_{a}$ does not have a sizeable effect except at very long distances or very high resolutions. Thus, it is natural that $\mu_{a}$ has not yet been observed in any experiment. Moreover, this term of the Hamiltonian probably has no direct astrophysical implications. \section{Acceleration Moment Effects Allowed by Symmetry} \label{sec-symm} At this point, it is not at all clear what sort of effects will occur on the length scale $x_{a}$. Fischbach et al.\ identify the term (\ref{eq-gterm}) as a standard spin-orbit coupling term, since \begin{equation} \mbox{{\boldmath $\sigma$}}\cdot({\bf g}\times{\bf p})=-\frac{\Phi c^{2}} {r^{2}}\,\mbox{{\boldmath $\sigma$}}\cdot({\bf x}\times{\bf p})=-\frac{2\Phi c^{2}}{\hbar r^{2}}\,{\bf L\cdot S}. \end{equation} This is correct in general, but it is not useful in the approximation of a uniform gravitational field ${\bf a}=-{\bf g}$, since the meaningful behavior of ${\bf L= x\times p}$ relies on ${\bf x}$ not remaining constant. We shall therefore work only with (\ref{eq-aterm}), which is clearly not a simple ${\bf L\cdot S}$ interaction. In fact, we shall find that the simple precession characteristic of ${\bf L\cdot S}$ interactions does not occur in this problem. Our arguments will be qualitative, based on the symmetries of the situation, so the numerical difference between (\ref{eq-aterm}) and (\ref{eq-gterm}) is not significant. As noted in~\cite{ref-maltschul}, $\mbox{{\boldmath $\chi$}}'$ does not affect the space-time trajectory of a neutron. However, $\mbox{{\boldmath $\chi$}}'$ may affect the spin vector, similar to the effect of a magnetic field, ${\bf B}$. If we express the direction of the spin vector as $(\theta,\phi)$ and apply the magnetic field ${\bf B}$ along the $z$-axis, $\phi$ precesses. However, the acceleration vector ${\bf a}$ possess a higher symmetry than ${\bf B}$, making the situation more complex. The symmetry point group of ${\bf B}$ is $SO(2)$, since the physics are unchanged by rotations about the $z$-axis. The acceleration ${\bf a}$ exhibits the same properties. However, the magnetic field is an axial vector, changing sign under reflection. Acceleration is a vector, so its symmetry point group is $O(2$). We will consider the action of a known acceleration ${\bf a}$, suitably parameterized. The operation of the acceleration on the spin vector $(\theta,\phi)$ must commute with the elements of $O(2)$. This means that $\phi$ must remain constant, since no rotation about the $z$-axis commutes with the reflections in $O(2)$. So the only rotations allowed are in the polar angle $\theta$. These rotations may not take $\theta$ to 0 or $\pi$, since $\phi$ is indeterminate at these poles, so the transformation would not be invertible. If $\theta$ is the only coordinate changing, it can not oscillate, since it is restricted to the interval $(0,\pi)$, and the transformation induces a rotation of definite direction at each point of this interval. Changes in $\theta$ must rotate the spin vector either toward or away from the $z$-axis, the direction of acceleration. Although the rate of this rotation may vary, it must satisfy \begin{equation} \dot{\theta}|_{\theta=\theta_{0}}=\dot{\theta}|_{\theta=\pi-\theta_{0}} \label{eq-theta} \end{equation} in order to be consistent with the action of the inverse transformation. The inverse, which would correspond to acceleration along the $-z$-axis, inverts $\dot{\theta}$. Symmetry dictates that if, at the angle $\theta_{0}$, the regular rotation was toward the $z$-axis, the inverse rotation at the angle $\pi-\theta_{0}$ must be toward the $-z$-axis at the same rate. Since the inverse transformation just reverses the rotation direction, the regular rotation at $\pi-\theta_{0}$ is the same as it is at $\theta_{0}$. Although we have not demonstrated this fact, it seems likely that the range of rotation of $\theta$ is the entire range $(0,\pi)$, with the sign of $\dot {\theta}$ the same everywhere. This would result in the spin vector continuously rotating towards or away from the direction of the acceleration, with the angular frequency symmetric about $\theta=\frac{\pi}{2}$ and gradually decreasing as $\theta$ approaches the pole. \section{Conclusions} The existence of $\mu_{a}$ does not ensure the validity of the equivalence principle. $\mbox{{\boldmath $\chi$}}'$ itself gives us no indication of the magnitude of the ``anomalous acceleration moment.'' However, $\mu_{a}$ does provide a convenient method for restoring the equivalence principle, if $\mu_{a}=-3$. This specific prediction gives us another avenue for analysis of both $\mu_{a}$ and the equivalence principle. The particular value of $\mu_{a}$ is a property of the quark operators under acceleration. By examining the problem from the QCD standpoint, it may be possible to determine $\mu_{a}$ directly, either confirming or rejecting the equivalence principle. If $\mu_{a}$ exists for the neutron, it almost certainly exists for the other neutral baryons. Whether it has an analogue for other neutral particles is not as clear. For higher-spin mesons, such as the $\rho(770)^{0}$, with spin 1, there may well be other ``anomalous acceleration moments.'' To see how such moments would enter the Hamiltonian, it is necessary to examine the acceleration generators for these particles. \acknowledgments The author wishes to thank Martin S. Altschul for his extensive and useful comments.
\section{Introduction.} In particle physics fields are given mass by assuming the existence of Higgs scalar fields. There are a number of unsatisfactory aspects to this procedure, for example: the Higgs scalar has not been experimentally found, also it is required to have a mass term of the wrong sign. An alternative procedure has been suggested, Roberts (1989) \cite{bi:mdr89} in which fluids rather than scalar fields are responsible for vector fields acquiring mass. The motivation for using fluids is a principle \cite{bi:mdr89} which states that there is only one concept of mass in physics; furthermore this concept of mass must be ultimately of gravitational origin. The picture envisaged can be thought of as occurring in four steps or stages, these words have unwanted temporal connotations so that we refer instead to four basic ingredients. The first ingredient is that the vector field is taken to exist, and it is assumed to have a {\bf primitive stress}. The second ingredient is that the vector field has statistical properties which produce an {\bf effective fluid} that couples to the primitive stress. The third ingredient is that the coupling between the fluid and the primitive stress produces a {\bf mass term}. The fourth ingredient is that the {\bf gravitational origin} of this is in some as yet undiscovered relationship between statistical mechanics and gravitational theory. Those of a more prosaic disposition can simply regard fluids as an alternative technical means for inducing mass, perhaps with application in the theory of superfluids, Isreal (1981) \cite{bi:isreal}. Scalar fields and fluids can be equated in several ways as shown for example in \cite{bi:mdr89}. Usually fluids exhibit more freedom as they are parameterised by $p, \mu, V_a$ however static scalar fields are equivalent to fluids with imaginary vector $V_a$ so that there are sometimes different qualitative properties, an example of this is the asymptotic properties, Roberts (1998) \cite{bi:mdr98}of the space-time. The technical methods in \cite{bi:mdr98}, uses what in section \ref{sec:II} is called the "already interacting" fluid. Here the technical procedure is to replace derivatives in the velocity decomposition of the fluid tangent vector $V_a$ by vector covariant derivatives. The resulting "covariantly interacting" fluid can be reduced to scalar electrodynamics, and then the standard symmetry breaking procedure applied. The covariantly interacting fluid generalizes scalar electrodynamics and the extra degrees of freedom provide the scope for other symmetry breaking mechanisms to be used. One mentioned here requires that only quantities of thermodynamic importance be retained in the mass breaking term, these thermodynamic quantities can then be fixed at a critical value such as the Bose condensation value. The drawback of this approach in its present form is that it necessitates the introduction of an ad hoc time interval; here this is taken to be the Planck time. It is found that the induced mass is negligible, i.e. well beneath the experimental limit of $10^{-50}Kg.$ \cite{bi:GN}. It is of interest to know if it could be demonstrated that the photon mass is exactly zero. The Proca equation \cite{bi:IZ} p.135 necessitates $m^2\nabla_aA^a=0$, so that the existence of photon mass fixes the gauge. The Bohm-Aharonov effect \cite{bi:BH} shows that the gauge must be chosen such that $A_a$ is continuous; it might be possible to devise a geometric configuration in which the continuity of $A_a$ requires a different gauge choice from $\nabla_aA^a=0$, thus giving $m=0$ from $m^2\nabla_aA^a=0$. In this paper section \ref{sec:II} briefly introduces the standard scalar electrodynamic symmetry breaking, section \ref{sec:III} introduces the velocity decomposition for the velocity tangent to a fluid, section \ref{sec:IV} applies the vector covariant derivative to this, section \ref{sec:V} mentions some possible ways in which the resulting fluid could break symmetry. The conventions used are: signature $-+++$, $";"$ signifies Christoffel covariant derivative, symmetrization is denoted by round brackets, e.g. $T_{(ab)}=(T_{ab}+T_{ba})/2$, anti-symmetrization is denoted by square brackets, e.g. $T_{[ab]}=(T_{ab}-T_{ba})/2$. \section{Scalar Electrodynamics.} \label{sec:II} The scalar electrodynamic Lagrangian \cite{bi:higgs}, \cite{bi:IZ}p.68, \cite{bi:HE}p.699 is \begin{equation} L_{sel}=-D_a\psi D^a\bar{\psi}-V(\psi\bar{ps})-\frac{1}{4}F^2, \label{eq:2.1} \end{equation}} %\indent where the covariant derivative is \begin{equation} D_a\psi=\partial_a\psi+ieA_a, \label{eq:2.2} \end{equation}} %\indent and $D_a\bar{\psi}=\bar{D_a\psi}$. The variation of the corresponding action with respect to $A_a, \psi$, and $\bar{ps}$ are given by \begin{eqnarray} \frac{\delta I}{\delta A_c}&=&F^{ab}_{..;b} +ie(\psi D^a\bar{\psi}-\bar{ps}D^a\psi)\nonumber\\ \frac{\delta I}{\delta \psi}&=&(D_aD^a-V')\bar{\psi},~~~V'=\frac{dV}{d(\psi\bar{\psi})}, \label{eq:2.3} \end{eqnarray} and its complex conjugate. Variations of the metric give the stress \begin{equation} T_{ab}=2D_{(a}\psi D_{b)}\bar{\psi}+F_{ac}F^{~c}_{b.}+g_{ab}L. \label{eq:2.4} \end{equation}} %\indent The complex scalar field can be put in "polar" form by defining \begin{equation} \psi= exp(i\nu), \label{eq:2.5} \end{equation}} %\indent giving the Lagrangian \begin{equation} L=\rho^2_a+({\mathcal D}_a\nu)^2-V(\rho^2)-\frac{1}{4}F^2, \label{eq:2.6} \end{equation}} %\indent where \begin{equation} {\mathcal D}_a\nu=\rho(\nu_a+eA_a). \label{eq:2.7} \end{equation}} %\indent The variations of the corresponding action with respect to $A_a ,\rho$, and $\nu$ are given by \begin{eqnarray} \frac{\delta I}{\delta A_a}&=&F^{ab}_{..;b}+2e\rho{\mathcal D}^a\nu,\nonumber\\ \frac{\delta I}{\delta \rho}&=&2\left(\Box+(\nu_a+eA_a)^2+V'\right),\nonumber\\ \frac{\delta I}{\delta \nu}&=&2\left(\Box\nu+eA^a_{,a;a}\right). \label{eq:2.8} \end{eqnarray} Variation of the metric gives the stress \begin{equation} T_{ab}=2\rho_a\rho_b +2{\mathcal D}_a\nu{\mathcal D}_b\nu+F_{ac}F^{~c}_{b.} +g_{ab}L. \label{eq:2.9} \end{equation}} %\indent Defining \begin{equation} B_a=A_a+\nu_a/e, \label{eq:2.10} \end{equation}} %\indent $\nu$ is absorbed to give Lagrangian \begin{equation} L=\rho^2_a+\rho^2e^2B_a^2-V(\rho^2)-\frac{1}{4}F^2, \label{eq:2.11} \end{equation}} %\indent which does not contain $\nu$; equation \ref{eq:2.10} is a gauge transformation when there are no discontinuities in $\nu$ i.e. $\nu_{;[ab]}=0$. The requirement that the corresponding quantum theory is renormizable restricts the potential to the form \begin{equation} V(\rho^2)=m^2\rho^2+\lambda\rho^4. \label{eq:2.12} \end{equation}} %\indent The ground state is when there is a minimum, for $m^2,\lambda>0$ this is $\rho=0$, but for $m^2<0,\lambda>0$ this is \begin{equation} \rho^2=\frac{-m^2}{2\lambda}=a^2, \label{eq:2.13} \end{equation}} %\indent thus the vacuum energy is \begin{equation} <0|\rho|0>=a. \label{eq:2.14} \end{equation}} %\indent To transform the Lagrangian \ref{eq:2.11} to take this into account substitute \begin{equation} \rho\rightarrow\rho'=\rho+a, \label{eq:2.15} \end{equation}} %\indent to give \begin{equation} L=\rho_a^2+(\rho+a)^2e^2B_a^2-V\left((\rho+a)^2\right)-\frac{1}{4}F^2. \label{eq:2.16} \end{equation}} %\indent Now apparently the vector field has a mass $m$ from the $a^2e^2B_a^2$ term it is given by $m=ae$. The cross term $2a\rho e^2B_a^2$ is ignored. \section{Velocity Potentials.} \label{sec:III} A Newtonian $3$-vector, can be decomposed: $\nu=\nabla\phi+\alpha\nabla\beta$, \cite{bi:clebsch} where $\phi, \alpha$, and $\beta$ are the Clebsch potentials. A particular case of Paff's theorem shows that a $4$-vector can be decomposed: $V_a=AB_a+CD_a$ where $A, B, C$, and $D$ are the potentials. The work of \cite{bi:lin} \cite{bi:SW} \cite{bi:schutz70} \cite{bi:schmid1} \cite{bi:schmid2} shows that a non-minimal decomposition \begin{equation} V_a=h^{-1}(\phi_a+\alpha\beta_a-\theta S_a),~~~V_aV^a_.=-1 \label{eq:3.1} \end{equation}} %\indent is more useful, because for an isentropic fluid all the potentials have evolution equations. $\theta$ is the thermasy of van Dantzig \cite{bi:vand}eq.4.9 it is usually defined by \begin{equation} d\theta=-kT~d\tau, \label{eq:3.2} \end{equation}} %\indent where $T$ is the temperature; also $h$ is the enthalpy and $S$ is the entropy. The three other potentials do not have a thermodynamic interpretation. A current vector can be defined by \begin{equation} W_a=hV_a, \label{eq:3.3} \end{equation}} %\indent c.f.\cite{bi:HE}p.69, because $W^a_{.;a}=0$ it is not conserved. The Christoffel derivative of $W_a$ can be in the usual manner \cite{bi:HE}p.82-3 \begin{equation} W_{a;b}=\omega_{ab}+\sigma_{ab}+\frac{1}{3}\Theta h_{ab}-\dot{W}_aW_b, \label{eq:3.4} \end{equation}} %\indent and then defining the vorticity tensor, vorticity vector, expansion tensor, expansion scalar, shear tensor, and acceleration vector in the usual manner, it is found that only the vorticity tensor, expansion scalar, and acceleration vector show any simplification, they are \begin{eqnarray} \omega _{ab}&\equiv&h^{~c}_{a.}h^{~a}_{b.}W_{[c;d]} =h^{~c}_{a.}h^{~a}_{b.}(\alpha_{[d}\beta_{c]}-\theta_{[d}S_{c]}),\nonumber\\ \Theta&\equiv&W^a_{.;a} =\Box\phi+\alpha\Box\beta-\theta\Box S+\alpha^a\beta_a-\theta^aS_a,\nonumber\\ \dot{W}_a&\equiv&V^bW_{a;b} =-TS_a+V^a(\phi_{ab}+\alpha\beta_{ab}-\theta S_{ab}), \label{eq:3.5} \end{eqnarray} respectively, where the projection tensor is given by \begin{equation} h_{ab}=g_{ab}-V_aV_b, \label{eq:3.6} \end{equation}} %\indent and the evolution equations \ref{eq:3.13} have been used in deriving the equation for the acceleration $\dot{W}_a$. The evolution equations can be derived from a Lagrangian; to do this it is necessary to assume both the equation \begin{equation} nh=p+\mu, \label{eq:3.7} \end{equation}} %\indent where $n$ is the particle number, $p$ is the pressure, and $\mu$ is the density, and also the first law of thermodynamics in the form \begin{equation} dp=n~dh-nT~dS. \label{eq:3.8} \end{equation}} %\indent Usually \cite{bi:hargreaves} the the Lagrangian of a fluid is taken to be the pressure, this is done here; but occasionally. e.g.\cite{bi:HE}p.69, other quantities are used. The action is \begin{equation} I=\int\sqrt{-g}p~dx. \label{eq:3.9} \end{equation}} %\indent Using \ref{eq:3.8} and then \ref{eq:3.1} shows that variations in the pressure depend on the velocity potentials. The variations are \begin{eqnarray} \frac{\delta I}{\delta g^{ab}}&=&-nhV_aV_b,\nonumber\\ \frac{\delta I}{\delta\phi}&=&-\sqrt{-g}(nV^a)_{;a},\nonumber\\ \frac{\delta I}{\delta\alpha}&=&-\sqrt{-g}n\beta^aV_a,\nonumber\\ \frac{\delta I}{\delta\beta}&=&-\sqrt{-g}(n\alpha V^a)_{;a},\nonumber\\ \frac{\delta I}{\delta\theta}&=&+\sqrt{-g}nS^aV_a,\nonumber\\ \frac{\delta I}{\delta S}&=&-\sqrt{-g}[(n\theta V^a)+nT]. \label{eq:3.10} \end{eqnarray} The first of these variations give the stress \begin{equation} T_{ab}=(p+\mu)V_aV_b+pg_{ab}. \label{eq:3.11} \end{equation}} %\indent The absolute derivative is given by \begin{equation} \dot{X}_{ab\ldots}=\frac{D}{d\tau}X_{ab\ldots}=V^cX_{ab\ldots;c}, \label{eq:3.12} \end{equation}} %\indent then the requirement that the second of the variations \ref{eq:3.10} vanishes is \begin{equation} \dot{n}+nV^a_{.;a}=0, \label{eq:3.13} \end{equation}} %\indent and this is just the conservation of particle number \cite{bi:stewart}eq.2.3. Using this the vanishing of the remaining variations gives \begin{equation} \dot{\alpha}=\dot{\beta}=\dot{S}=0,~~~\dot{\theta}= -T. \label{eq:3.14} \end{equation}} %\indent The third of these shows that the fluid is isentropic, the fourth is just \ref{eq:3.2} in another forms. Using \ref{eq:3.1} and \ref{eq:3.14} shows that \begin{equation} \dot{\phi}=-h. \label{eq:3.15} \end{equation}} %\indent The Bianchi identities give the fluid conservation equations \begin{eqnarray} -V_aT^{ab}_{..;b}&=&\dot{\mu}+(p+\mu)V^a_{.;a},\nonumber\\ h_{ab}T^{bc}_{..;c}&=&(p+\mu)V^bV_{a;b}+h^{~b}_{a.}p_{;b}, \label{eq:3.16} \end{eqnarray} where the projection tensor is given by \ref{eq:3.5}; these equations do not immediately occur as a result of varying the Lagrangian, but only by applying the Bianchi identities to \ref{eq:3.11}. The first law of thermodynamics has been assumed \ref{eq:3.5}. In the present case the second law is obeyed as an equality as in \ref{eq:3.14} $\dot{S}=0$. The situation is more complex in the next section. Anticipating some of the problems and how to approach them - In particle physics there is not always invariance under time and space reflections. The above involves no inequalities: there are no assumed energy inequalities, and the fluid is isentropic (in equilibrium with $\dot{S}=0$) rather than $\dot{S}>0$. Recall that a vector is future pointing iff $V_t>0$; for \ref{eq:3.1} $hV_t=\phi_t+\alpha\beta_t-\theta S_t$; ignoring $\phi, \alpha$, and $\beta$, $V_a$ is past pointing when $\theta S_t >0$. This can be reversed by either defining a new vector $V'_a=-V_a$ or a new time coordinate $t'=-t$. \section{The "Covariantly Interacting" Fluid.} \label{sec:IV} There are several ways, four of which are mentioned here, of introducing an interacting fluid and vector field, with the intention of breaking the vector fields symmetries. In the first method a "plasma interacting" fluid is produced by generalizing the treatment of \cite{bi:SW}$\verb+#+$7. In the second method, the fluid is "already interacting", the stress \ref{eq:3.10} is directly equated with the stress calculated from \ref{eq:2.11}; this gives $-2(\rho_a\rho_b+\rho^2e^2B_aB_b)=(\mu+p)V_aV_b$ and it is impossible to proceed with one real fluid; this method is similar to the method discussed in \cite{bi:mdr89}. The third method is the "traditionally interacting" fluid \cite{bi:HE}p.70, this is produced by adding a term $L_I=-\frac{1}{2}V_aA^a$ to the Lagrangian; this method is of no use for present purposes as there is only a single term $A_a$ in the Lagrangian and stress, and for symmetry breaking products $A_a A_B$ are required. The fourth method is the "covariantly interacting" fluid; this is a produced by simply applying the covariant derivative \ref{eq:2.2} to the vector \ref{eq:3.1} to produce \begin{equation} V_a=h^{-1}\left(\phi_a+\alpha\beta_a-\theta S_a+ie(\phi+\alpha\beta-\theta S)A_a\right). \label{eq:4.1} \end{equation}} %\indent It is required that \begin{equation} -1=g^{ab}_{..}V_{(a}\bar{V}_{b)}, \label{eq:4.2} \end{equation}} %\indent and also that the projection tensor is \begin{equation} h_{ab}=g_{ab}-V_{(a}\bar{V}_{b)}, \label{eq:4.3} \end{equation}} %\indent this ensures that most quantities are real. After using $W_a=hV_a$, \ref{eq:4.2} becomes \begin{equation} h=-g^{ab}_{..}W_{(a}\bar{V}_{b)}, \label{eq:4.4} \end{equation}} %\indent then via the first law of thermodynamics \ref{eq:3.7} the pressure and the Lagrangian are real; however the expansion, shear, \ldots etc. are complex, this can be verified by direct computation or by noting from \ref{eq:3.4} that as $W_a$ is complex the quantities involved in this decomposition should also be complex. The evolution equations are derived in a similar manner to the derivation in the last section. The action is taken to be \ref{eq:3.9} and the first law of thermodynamics is assumed with the enthalpy $h$ now given by \ref{eq:4.4}. The variations of the action are \begin{eqnarray} \frac{\delta I}{\delta g^{ab}_{..}}&=&-nhV_{(a}\bar{V}_{b)},\nonumber\\ \frac{\delta I}{\delta A^a_.}&=&-n\sqrt{-g}\frac{e'^2}{h^2}A_a, ~~~e'^2=e^2(\phi+\alpha\beta-\theta S),\nonumber\\ \frac{\delta I}{\delta \phi}&=&-\sqrt{-g}[(n\Re(V^a_.))+e'^2A_a^2n],\nonumber\\ \frac{\delta I}{\delta \alpha}&=&-n\sqrt{-g}[\beta_a\Re(V^a_.)+e'^2A_a^2\beta],\nonumber\\ \frac{\delta I}{\delta \beta}&=&-\sqrt{-g}[(n\alpha\Re(V^a_.)_{;a} +e'^2A_a^2n\alpha],\nonumber\\ \frac{\delta I}{\delta \theta}&=&+n\sqrt{-g}[S_a\Re(V^a_.)+e'^2A_a^2S],\nonumber\\ \frac{\delta I}{\delta S}&=&+\sqrt{-g}[(n\theta\Re(V^a_.))_{;a}+e'^2A_a^2\theta n-nT], \label{eq:4.5} \end{eqnarray} \ref{eq:4.4} and the vanishing of these give \begin{equation} -h=[\phi_a\Re(V^a_.)+e'^2A_a^2\phi]. \label{eq:4.6} \end{equation}} %\indent The stress and Maxwell equation are \begin{eqnarray} T_{ab}&=&(p+\mu)V_{(a}\bar{V}_{b)}+p~g_{ab},\nonumber\\ F^{ab}_{..}&+&\frac{e}{h^2}(\phi+\alpha\beta-\theta S)A^a_.= 0. \label{eq:4.7} \end{eqnarray} Introducing the notation \begin{eqnarray} \dot{X}_{ab\ldots}&=&\frac{DX_{ab\ldots}}{d\tau} =\Re(V^c_.)X_{ab\ldots;c},\nonumber\\ \hat{X}_{ab\ldots}&=&\left(\frac{D}{d\tau}+e'^2A_a^2\right)X_{ab\ldots},\nonumber\\ \label{eq:4.8} \end{eqnarray} \ref{eq:4.6} and the vanishing of the variations \ref{eq:4.5} can be written in the simple form \begin{equation} \dot{\alpha}=\hat{\beta}=\hat{S}=0,~~~\hat{\phi}=-h,~~~\dot{\theta}=T,~~~\hat{n}=n\Theta\ne0. \label{eq:4.9} \end{equation}} %\indent Changing the sign and index of the last term produces the changes in the table. The fluid has $n\ne=0$ implying that particle number is not conserved, this appears unavoidable. Our choice has $\theta=T$ so that the standard expression for thermasy is recovered; also $\hat{S}=0$ implies that the second law of thermodynamics is obeyed when $\theta S>\phi+\alpha\beta$.\newpage Table: The Choice for the Thermodynamic Term in the Velocity Vector. \begin{tabular}{||l|l|l|l||} \hline\hline $ -\theta_a S $ & $ +\theta S_a $ & $ +\theta_a S $ & $ -\theta S_a$\\ \hline $\hat{\phi}=ST-h$ & $ \hat{\phi}=-h$& $\hat{\phi}=ST-h$ & $\hat{\phi}=-h$\\ $ \dot{S}=0 $ & $ \hat{S}=0 $ & $\dot{S}=0 $ & $\hat{S}=0$\\ $ \hat{\theta}=-T $ & $ \dot{\theta}=-T$& $\hat{\theta}=-T$ & $\dot{\phi}=-T$\\ \hline\hline \end{tabular} \section{Symmetry Breaking.} \label{sec:V} The straightforward way of proceeding to break symmetry is to note that the covariantly interacting fluid of the previous section contains scalar electrodynamics as a special case. There are several choices of the parameters by which scalar electrodynamics can be recovered, an example is \begin{equation} \alpha=\beta=\theta=S=\partial_a h=0,~~~\phi=\sqrt{2}h\psi,~~~p=l_{sel},~~~\mu=1-p, \label{eq:5.1} \end{equation}} %\indent in the gauge ref{eq:2.10}. Thus the standard way of breaking symmetry can then be applied, with the aesthetic difference the fundamental cause of breaking is a fluid not a field. There is the possibility of other velocity potentials such as $\alpha$ being non-vanishing operators leading to modifications of the standard treatment; the velocity potentials themselves have been subject to quantization for a fluid coupled to gravity in the ADM formalism, see for example \cite{bi:schutz71} \cite{bi:DM} \cite{bi:LR}. The Proca equation \cite{bi:IZ}p.135 is \begin{equation} F^{ab}_{..}+m^2eA^a_.=0, \label{eq:5.2} \end{equation}} %\indent comparing with (\ref{eq:4.8}, the covariantly interacting fluid has a mass \begin{equation} m^2=h^{-2}(\phi+\alpha\beta-\theta S)^2. \label{eq:5.3} \end{equation}} %\indent Choosing only to retain thermodynamic quantities $m^2=h^{-2}\theta^2S^2$, and the thermasy $\theta$ needs to be evaluated. The tempreture can be taken to be independent of the proper time $\tau$, but the mass in \ref{eq:5.3} still depends on proper time; to proceed it is necessary to introduce an artificial time interval and the Plank time is chosen, thus \begin{equation} -\theta= k\int^{ta_{PL}}_0~T~d\tau = kT\int^{\tau_{PL}}_0~d\tau =kT~|^{\tau_{PL}}_0 =\sqrt{\frac{G\hbar}{c^5}}kT. \label{eq:5.4} \end{equation}} %\indent Restoring constants and substituting into \ref{eq:5.3} gives \begin{equation} m=\sqrt{\frac{c\hbar}{G}}k\frac{TS}{h}\ap3.10^{-31}\frac{TS}{h}~ Kg. \label{eq:5.5} \end{equation}} %\indent For $m$ to be constant it is necessary to evaluate $TS/h$ at the critical point and the Bose condensation point is a choice. This choice terms out to give zero mass because Bose condensation requires that the lowest state has zero kinetic energy, as there is no extra energy there is no extra heat content and hence no entropy, there is no disorder because all the particles are in the same state, c.f.\cite{bi:kahanna}p.78, no entropy implies that there is no induced mass. The Bose condensation point is also unsatisfactory choice because strictly speaking $N/V$ is not required to have a given value for radiation in a cavity for a vector fields, c.f.\cite{bi:jackson}prob.16.2.
\section{#1}} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \textwidth 170mm \textheight 230mm \topmargin -15mm \oddsidemargin -0.60 cm \evensidemargin -0.60 cm \parindent 1 true cm \newcommand{\lbl}[1]{\label{eq:#1}} \newcommand{ \rf}[1]{(\ref{eq:#1})} \newcommand{\vs}[1]{\rule[- #1 mm]{0mm}{#1 mm}} \newskip\humongous \humongous=0pt plus 1000pt minus 1000pt \def\mathsurround=0pt{\mathsurround=0pt} \def\eqalign#1{\,\vcenter{\openup1\jot \mathsurround=0pt \ialign{\strut \hfil$\displaystyle{##}$&$ \displaystyle{{}##}$\hfil\crcr#1\crcr}}\,} \newif\ifdtup \def\panorama{\global\dtuptrue \openup1\jot \mathsurround=0pt \everycr{\noalign{\ifdtup \global\dtupfalse \vskip-\lineskiplimit \vskip\normallineskiplimit \else \penalty\interdisplaylinepenalty \fi}}} \def\eqalignno#1{\panorama \tabskip=\humongous \halign to\displaywidth{\hfil$\displaystyle{##}$ \tabskip=0pt&$\displaystyle{{}##}$\hfil \tabskip=\humongous&\llap{$##$}\tabskip=0pt \crcr#1\crcr}} \def{\widehat m}{{\widehat m}} \newcommand{\vs{2}\begin{equation}}{\vs{2}\begin{equation}} \newcommand{\\[2mm]\end{equation}}{\\[2mm]\end{equation}} \newcommand{\begin{eqnarray}}{\begin{eqnarray}} \newcommand{\end{eqnarray}}{\end{eqnarray}} \newcommand{\stackrel{\displaystyle <}{{}_{\displaystyle\sim}}}{% \mathrel{% \setbox0=\hbox{$<$} \raise0.6ex\copy0\kern-\wd0 \lower0.65ex\hbox{$\sim$} }} \newcommand{\stackrel{\displaystyle >}{{}_{\displaystyle\sim}}}{% \mathrel{% \setbox0=\hbox{$>$} \raise0.6ex\copy0\kern-\wd0 \lower0.65ex\hbox{$\sim$} }} \defM_\pi{M_\pi} \defF_\pi{F_\pi} \newcommand{\cpm}[1]{[#1]_\pm} \newcommand{\cp}[1]{[#1]_+} \newcommand{\cm}[1]{[#1]_-} \newcommand{\alpha}{\alpha} \newcommand{\beta}{\beta} \newcommand{\gamma}{\gamma} \newcommand{\delta}{\delta} \newcommand{\rho}{\rho} \newcommand{\sigma}{\sigma} \newcommand{\tau}{\tau} \newcommand{\eta}{\eta} \newcommand{\NP}[1]{Nucl.\ Phys.\ {\bf #1}} \newcommand{\RMP}[1]{Rev.\ of Mod.\ Phys.\ {\bf #1}} \newcommand{\PL}[1]{Phys.\ Lett.\ {\bf #1}} \newcommand{\NC}[1]{Nuovo Cimento {\bf #1}} \newcommand{\AN}[1]{Ann. Phys. {\bf #1}} \newcommand{\CMP}[1]{Comm.\ Math.\ Phys.\ {\bf #1}} \newcommand{\PR}[1]{Phys.\ Rep.\ {\bf #1}} \newcommand{\PRev}[1]{Phys.\ Rev.\ {\bf #1}} \newcommand{\PRL}[1]{Phys.\ Rev.\ Lett.\ {\bf #1}} \newcommand{\MPL}[1]{Mod.\ Phys.\ Lett.\ {\bf #1}} \newcommand{\BLMS}[1]{Bull.\ London Math.\ Soc.\ {\bf #1}} \begin{document} \renewcommand{\thefootnote}{\fnsymbol{footnote}} \setcounter{equation}{0} \setcounter{subsection}{0} \setcounter{table}{0} \setcounter{figure}{0} \begin{titlepage} \begin{flushright} CPT-98/P.3716\\ LU TP 98/23\\ ZU-TH 21/98\\ \today\\ \end{flushright} \begin{center} \begin{bf} {\Large \bf Low--energy photon--photon fusion into three pions \\[0.5cm] in Generalized Chiral Perturbation Theory \footnote{Work supported in part by TMR, EC--Contract No. ERBFMRX--CT980169} } \\[2cm] \end{bf} {\large Ll. Ametller$^a$, J. Kambor$^b$, M. Knecht$^c$ and P. Talavera$^{d}$} \\[1cm] $^a$ Dept. de F{\'\i}sica i Enginyeria Nuclear, UPC, E-$08034$ Barcelona, Spain.\\[.1cm] $^b$ Institut f\"ur Theoretische Physik, Universit\"at Z\"urich, CH-8057 Z\"urich, Switzerland.\\[.1cm] $^c$ Centre de physique Th\'eorique, CNRS--Luminy, Case 90 F--13288 Marseille Cedex 9, France.\\[.1cm] $^d$ Dept. of Theoretical Physics, University of Lund, S\"olvegatan 14A, S-22362 Lund, Sweden. \\[3cm] {\bf PACS:}~11.30.Rd, 12.39.Fe, 13.75.-n, 14.70.Bh, 13.60.Le\\[0.2mm] {\bf Keywords:} \begin{minipage}[t]{9.5cm} Chiral Symmetry, Chiral Perturbation Theory, Scattering at low energy, Photon-photon Physics, Meson production.\end{minipage} \end{center} \vfill \begin{abstract} The processes $\gamma\gamma\to\pi^0\pi^0\pi^0$ and $\gamma\gamma\to \pi^+ \pi^- \pi^0$ are considered in Generalized Chiral Perturbation theory, in view of their potential sensitivity to the mechanism of spontaneous breaking of chiral symmetry and to various counterterms. The amplitudes are computed up to order ${\cal O}({\mbox p}^6)$. The event production rates are estimated for the Daphne $\phi$--Factory and for a future $\tau$--Charm Factory. \end{abstract} \vfill \end{titlepage} \setcounter{footnote}{0} \renewcommand{\thefootnote}{\arabic{footnote}} \section{\bf Introduction} In the limit where the masses of the lightest quark flavours $u,d$ and $s$ are set to zero, the QCD lagrangian becomes invariant under a chiral $SU(3)_{\mbox{\scriptsize L}}\times SU(3)_{\mbox{\scriptsize R}}$ global symmetry. This symmetry is not reproduced by the hadronic spectrum, and must therefore be spontaneously broken towards the diagonal $SU(3)_{\rm V}$ subgroup. Actually, this spontaneous breaking of chiral symmetry can be shown to follow from very general properties of QCD \cite{VaWi,tHo,Presk}. However, besides the existence of eight massless pseudoscalar states coupling to the eight conserved axial currents, nothing is known from ``first principles'' about the actual mechanism of spontaneous breaking of chiral symmetry. The widespread belief in that matter is that it proceeds through the formation of a strong quark--antiquark condensate, $<{\bar q}q>\sim -(250\,{\mbox MeV})^3$, where $<{\bar q}q>$ denotes the single flavour quark--antiquark condensate in the $SU(2)$ chiral limit, \vs{2}\begin{equation} <{\bar q}q> \,=\, <{\bar u}u>\vert_{m_u=m_d=0}\,=\, <{\bar d}d>\vert_{m_u=m_d=0}\ . \\[2mm]\end{equation} In particular, this means that once the light quark masses $m_u$, $m_d$ and $m_s$ are turned on, the mass of the pion is assumed to be dominated by the contribution linear in quark masses \cite{gor}, \vs{2}\begin{equation} -\frac{2{\widehat m} <{\bar q}q>}{F^2M_{\pi}^2}\sim 1\ ,\ \ {\widehat m} \,=\,\frac{m_u + m_d}{2}\ ,\lbl{stdcond} \\[2mm]\end{equation} where $F$ stands for the pion decay constant $F_\pi$ (the normalization we use corresponds to the numerical value $F_\pi$ = 92.4 MeV \cite{PDG}) in the same two--flavour chiral limit, \vs{2}\begin{equation} F = F_{\pi}\vert_{m_u=m_d=0}. \\[2mm]\end{equation} However, our present theoretical knowledge of non--perturbative aspects in QCD does not exclude a picture where the condensate would be much smaller, say $<{\bar q}q>\sim -(100\,{\mbox MeV})^3$, or even vanishing. How the latter possibility may arise in QCD has been discussed recently~\cite{stern98,stern97} in terms of spectral properties of the Dirac operator. On the other hand, it has also been suggested \cite{KKS98} that a strictly vanishing condensate in the chiral limit could be excluded by an inequality \cite{comellas95} between the correlator $<A_{\mu}(x)A^{\mu}(0)>$ of two axial currents and the two--point function $<P(x)P(0)>$ of the pseudoscalar quark bilinear density $P(x)\equiv ({\bar q}i\gamma_5q)(x)$. This inequality, however, has only been established so far for bare quantities, {\it i.e.} in the presence of an ultraviolet cut--off $\Lambda_{\mbox{\scriptsize UV}}$. As the cut--off is removed, the pseudoscalar density $P(x)$ needs to be renormalized, whereas the (partially) conserved axial current $A_{\mu}(x)$ remains unaffected. Stricktly speaking, the claim of Ref.~\cite{KKS98} is that $<{\bar q}q>(\Lambda_{\mbox{\scriptsize UV}})\neq 0$ for finite $\Lambda_{\mbox{\scriptsize UV}}$, which does not, {\it a priori}, exclude the possibility of a vanishing condensate in the limit $\Lambda_{\mbox{\scriptsize UV}}\to\infty$ \cite{stern98}. Finally, Ref! .~\cite{kneur} provides an example of a different approach which, within a variational framework, leads to a non--vanishing, but nevertheless small, value of $<{\bar q}q>$. Clearly, in order to settle the question of the mechanism of spontaneous chiral symmetry breaking on a purely theoretical level, a major breakthrough in our understanding of non--perturbative aspects of confining gauge theories is required. Instead, one may try a phenomenological approach, and look for experimental observables which could provide the relevant information. In this respect, low--energy $\pi$--$\pi$ scattering in the S--wave has been put forward as a process particularly sensitive to the size of the condensate \cite{pionscat,pipipaper1} (for recent discussions, see {\it e.g.} \cite{pionrev,stern97}). This can be most conveniently seen in the framework of the effective lagrangian \cite{weinberg78,GL1,GL2}. In order to incorporate the possibility of a small condensate, the usual counting has however to be modified. As explained in Refs.~\cite{pionscat,gchpt}, a consistent expansion scheme, usually refered to as Generalized Chiral Perturbation Theory (GChPT), is obtained by taking (from now on, we restrict ourselves to the case of two massless flavours only), \vs{2}\begin{equation} m_u,m_d\sim{\cal O}({\mbox p})\ ,\ \ B\sim{\cal O}({\mbox p})\ ,\lbl{gencount} \\[2mm]\end{equation} with p being a generic momentum, much smaller than the typical hadronic scale $\Lambda_H\sim 1\,{\mbox GeV}$, and \vs{2}\begin{equation} B\equiv -\frac{<{\bar q}q>}{F^2}\ . \\[2mm]\end{equation} With this counting, the effective lagrangian at lowest order, which consists of all chiral invariant terms of order ${\cal O}({\mbox p}^2)$, reads \cite{pionscat} \begin{eqnarray} {\tilde{\cal L}}^{(2)}&=&{1\over 4}F^2 \left\{\langle D_\mu U^+D^\mu U\rangle +2B\langle U^+\chi+\chi^+U\rangle \right.\nonumber\\ &&\qquad + A\langle (U^+\chi)^2+(\chi^+U)^2\rangle + Z^P\langle U^+\chi-\chi^+U\rangle ^2 \lbl{L2}\\ &&\qquad +\left. h_0\langle \chi^+\chi\rangle + h_1({\mbox{det}}\chi + {\mbox{det}}\chi^+)\right\}\ .\nonumber \end{eqnarray} The matrix $U$ collects the pion fields (throughout, we adopt the Condon and Shortley phase convention), \vs{2}\begin{equation} U\,=\, e^{i\phi/F}\ ,\ \ \phi = \left( \begin{array}{cc} \pi^0 & -\sqrt{2}\pi^+\\ \sqrt{2}\pi^- & -\pi^0 \end{array}\right)\ .\lbl{Ufield} \\[2mm]\end{equation} The notation is as in Refs.~\cite{GL1,GL2}, except that $\chi$, the quantity that contains the scalar and pseudoscalar sources, is defined without the usual factor $2B$, \vs{2}\begin{equation} \chi={s}+i{ p}={\cal M}+\cdots \ ,\, {\cal M}={\mbox{diag}}(m_u,m_d) \ .\lbl{chi} \\[2mm]\end{equation} The covariant derivative contains the external vector and axial sources, \vs{2}\begin{equation} D_{\mu}U\ =\ \partial_{\mu}U-i[v_{\mu},U]-i\{a_{\mu},U\}\ . \\[2mm]\end{equation} This is to be contrasted with Standard Chiral Perturbation Theory (SChPT)~\cite{GL1,GL2}, which assumes a large condensate, and hence the counting rule $m_u,m_d\sim{\cal O}({\mbox p}^2)$, $B\sim\Lambda_H$, so that only the first two terms on the right--hand side of Eq.~\rf{L2} are taken into account at leading order. It easily follows from~\rf{L2} that the lowest order expression of the pion mass is now given by \footnote{We neglect the mass difference $m_u -m_d$ and set $m_u=m_d={\widehat m}$.} \vs{2}\begin{equation} M_{\pi}^2\,=\,2{\widehat m} B \,+\, 4{\widehat m}^2A \,+\,\cdots\ ,\lbl{Mpilead} \\[2mm]\end{equation} where the ellipsis stands for higher order corrections. At the same level of approximation in the chiral expansion, the $\pi$--$\pi$ scattering amplitude can be written as~\cite{pionscat} \vs{2}\begin{equation} A(s\vert t,u)\,=\,\beta\frac{(s-\frac{4}{3}M_{\pi}^2)}{F^2}\,+\, \alpha\frac{M_{\pi}^2}{3F^2}\,+\,\cdots\ ,\lbl{Amplead} \\[2mm]\end{equation} with $\beta=1+{\cal O}({\widehat m})$, while at this order the parameter $\alpha$ is directly related to the $<{\bar q}q>$ condensate through \vs{2}\begin{equation} \alpha = 4 - 3\bigg({{2{\widehat m} B}\over{M_\pi^2}}\bigg)\,+\,{\cal O}({\widehat m})\ . \lbl{alphalead} \\[2mm]\end{equation} The case~\rf{stdcond} of a strong condensate corresponds to $\alpha =1$, whereas the extreme limit where the condensate would vanish yields $\alpha = 4$. Notice that the above expression for $A(s\vert t,u)$ is not affected by the additional terms in \rf{L2} and reproduces the result first obtained by Weinberg~\cite{wein66}, $A(s\vert t,u)\,=\,(s-2{\widehat m} B)/F^2\,+\,\cdots$. The difference between the standard case and deviations from it lies here only in the leading--order expression~\rf{Mpilead} of the pion mass and its relation to the condensate. Higher orders in the chiral expansion will modify the simple expression \rf{Amplead}, but the correlation between low--energy $\pi$--$\pi$ scattering and the value of the ratio $2{\widehat m} B/M_{\pi}^2$ subsists, and can be studied in a controled way within the generalized chiral expansion \cite{pipipaper1}. Available data on low--energy $\pi$--$\pi$ phases, which are dominated by the Geneva--Saclay $K_{\ell 4}$ experiment~\cite{ross77}, do however not possess the required accuracy in order to distinguish between the different alternatives at present. Forthcoming experiments, such as new $K_{\ell 4}$ experiments, conducted by the BNL865 collaboration \cite{b! nl865} or planed at the Daphne $\phi$--factory \cite{daphne}, and the DIRAC experiment at CERN \cite{dirac}, represent promissing prospects in this direction. In order to create a (hopefully convergent) set of evidences {\it pro} or {\it contra} a specific picture of spontaneous breaking of chiral symmetry, it remains however important to explore other possibilities, and to find other processes which, at the theoretical level, can be shown to exhibit a reasonably strong dependence on the value of the condensate. The present work was motivated by the above consideration and the following observation. A straightforward re--analysis in GChPT of the existing SChPT calculations~\cite{adler,bos,pere96} at lowest order shows that the two amplitudes for the production of three pions in low--energy photon--photon collisions involve the parameter $\alpha$ already at tree level. In the case of $\gamma\gamma\to\pi^0\pi^0\pi^0$, the amplitude reads \vs{2}\begin{equation} {\cal A}^{N} = {{e^2}\over{4\pi^2F_\pi^3}}\epsilon^{\mu\nu\alpha\beta}k_\mu\epsilon_\nu k_\alpha'\epsilon_\beta' \,{{\alpha M_\pi^2}\over{s-M_\pi^2}}\,+\,\cdots\ , \\[2mm]\end{equation} while for $\gamma\gamma\to \pi^+\pi^-\pi^0$ we find \begin{eqnarray} {\cal A}^{C} &=& {{e^2}\over{4\pi^2F_\pi^3}}\epsilon_{\mu\nu\alpha\beta}k^\mu\epsilon^\nu k'^\alpha\epsilon'^\beta\,\bigg(1-{{(p_++p_-)^2-M_\pi^2+{1\over 3}(\alpha -1)M_\pi^2}\over {s-M_\pi^2}}\bigg) \nonumber\\ && +{{e^2}\over{4\pi^2F_\pi^3}} \epsilon_{\mu \nu \alpha \beta} k'^\mu \epsilon'^\nu p_0^{\alpha} \Bigl(-\epsilon^\beta+ {\epsilon \cdot p_{-}\over k\cdot p_{-}}p_+^{\beta} + {\epsilon \cdot p_{+}\over k\cdot p_{+}}p_-^{\beta}\Bigr) \nonumber\\ && +{{e^2}\over{4\pi^2F_\pi^3}} \epsilon_{\mu \nu \alpha \beta} k^\mu \epsilon^\nu p_0^{\alpha} \Bigl(-\epsilon'^\beta + {\epsilon' \cdot p_{-}\over k'\cdot p_{-}}p_+^{\beta} + {\epsilon' \cdot p_{+}\over k'\cdot p_{+}} p_-^\beta\Bigr)\,+\,\cdots\ . \end{eqnarray} In these expressions, $k$ and $k'$ ( $\epsilon$ and $\epsilon'$) denote the momenta (polarizations) of the two photons, while $p_+$, $p_-$ and $p_0$ are the pion momenta. The ellipses stand for higher order corrections which will be considered later. Diagrammatically, the origin of the dependence on $\alpha$ lies in the presence of one--pion reducible contributions to the two amplitudes, see Fig.~1 below. \indent \centerline{\psfig{figure=fig1.gg3pi.ps,height=3.5cm}} \noindent {\bf Figure 1} : {\it The lowest order contributions to the $\gamma\gamma\to\pi\pi\pi$ amplitudes. The $\alpha$--dependence in ${\cal A}^{N}$ and ${\cal A}^{C}$ comes from the vertex for (virtual) $\pi$--$\pi$ scattering in the first graph, which is the only one to contribute in the case of the neutral amplitude.} \indent \noindent Therefore, the leading--order neutral amplitude ${\cal A}^N$ increases as the condensate decreases. For a strictly vanishing condensate ($\alpha =4$), the cross section at low energies is thus enhanced by a factor 16 as compared to the standard case of a strong condensate ($\alpha =1$) ! In the charged case, the situation is less favourable. At threshold, the average over the photon polarizations of the modulus squared of the amplitude is still proportional to the square of $\alpha$, \vs{2}\begin{equation} {1\over 4}\sum_{\mbox{\scriptsize pol}}\vert{\cal A}^C \vert^2_{\mbox{\scriptsize thr}} ={1\over 4}\,\bigg({{e^2}\over{4\pi^2F_\pi^3}}\bigg)^2\, {{9M_\pi^4}\over 128}\,\alpha^2\ , \\[2mm]\end{equation} but the sensitivity on $\alpha$ of the corresponding total {\it cross section} $\sigma^{C}(s,\alpha)$ rapidly decreases with increasing energy. For instance, the ratio $\sigma^C(s,\alpha=3)/\sigma^C(s,\alpha=1)$ is equal to 4.48 at $\sqrt{s}=450$ MeV, {\it i.e.} just above threshold, but drops to 1.68 at $\sqrt{s}=500$ MeV and becomes less than 1.10 at $\sqrt{s}\stackrel{\displaystyle >}{{}_{\displaystyle\sim}} 600$ MeV. For the case $\alpha =2$, the corresponding ratio is equal to 2.21 at $\sqrt{s}=450$ MeV, but the effect is less than 25\% at $\sqrt{s}=500$ MeV, while it barely reaches a few percent at $\sqrt{s}\stackrel{\displaystyle >}{{}_{\displaystyle\sim}} 600$. Thus, the interference between the two kinematical structures contributing to the amplitude ${\cal A}^C$, which is responsible, at low energies, for the suppression of the cross section in the charged channel as compared to the cross section in the neutral channel~\cite{pere96}, also washes out the dependence on the value of the $<{\bar q}q>$ condensate as soon as one leaves the threshold region. Considering definite polarization configurations for the two photons does not improve the situation~: For parallel polarizations of the two photons, the part of ${\cal A}^C$ which is sensitive to $\alpha$ does not contribute to the cross section, whereas for polarizations taken along orthogonal axes, the same interference effect is again fully at work. Furthermore, both cross sections rapidly rise above threshold, so that the effect of higher orders also needs to be investigated. The purpose of this paper is precisely to investigate these higher order effects at the one--loop level in GChPT. In fact, at next--to--leading order new tensorial structures appear both in the neut! ral and in the charged amplitude \cite{pere96}, whereas the pion loops provide additional sources of dependence with respect to $\alpha$, so that their behaviour with respect to changes in the value of the condensate could be modified to some extent. Accordingly, the rest of the paper is organized as follows~: The general kinematical structure of the two amplitudes ${\cal A}^N$ and ${\cal A}^C$, as well as their properties under isospin symmetry, are the subject of Section 2. The construction and the renormalization of the effective lagrangian of GChPT at order one--loop in the two--flavour case are treated in Section 3. Section 4 is devoted to the actual calculation of the two amplitudes to one--loop precision. The counterterms which are involved in these expressions and several numerical results are discussed in Section 5. A summary and conclusions are presented in Section 6. Details on the evaluation of some counterterms have been gathered in an Appendix. \indent \setcounter{equation}{0} \section{Kinematics and Isospin Symmetry} \setcounter{equation}{0} The amplitudes ${\cal A}^N$ and ${\cal A}^C$ for the processes \begin{eqnarray} \gamma(k)\ \gamma(k')\ &\to & \pi^0(p_1)\ \pi^0(p_2)\ \pi^0(p_3)\ , \nonumber\\ \gamma(k)\ \gamma(k')\ &\to & \pi^+(p_+)\ \pi^-(p_-)\ \pi^0(p_0)\ ,\lbl{proc} \end{eqnarray} are obtained from the matrix elements \begin{eqnarray} <\,\pi^0(p_1)\pi^0(p_2)\pi^0(p_3)\,{\mbox{out}}\,\vert\, \gamma(k,\epsilon)\gamma(k',\epsilon')\,{\mbox{in}}\,>&=& i(2\pi)^4\delta^4(P_f-P_i){\cal A}^N\ ,\nonumber\\ <\,\pi^+(p_+)\pi^-(p_-)\pi^0(p_0)\,{\mbox{out}}\,\vert\, \gamma(k,\epsilon)\gamma(k',\epsilon')\,{\mbox{in}}\,>&=& i(2\pi)^4\delta^4(P_f-P_i){\cal A}^C\ , \lbl{mat} \end{eqnarray} respectively, with \begin{eqnarray} && {\cal A}^N(k,\epsilon;k',\epsilon';p_1,p_2,p_3) \,=\, \nonumber\\ && \qquad ie^2\epsilon_{\mu}(k)\epsilon_{\nu}'(k') \int d^4x e^{-ik\cdot x} <\,\pi^0(p_1)\pi^0(p_2)\pi^0(p_3)\,{\mbox out}\,\vert\, T\{j^{\mu}(x)j^{\nu}(0)\}\,\vert\,\Omega\,>\ , \nonumber\\ && {\cal A}^C(k,\epsilon;k',\epsilon';p_0,p_+,p_-) \,=\, \nonumber\\ && \qquad ie^2\epsilon_{\mu}(k)\epsilon_{\nu}'(k') \int d^4x e^{-ik\cdot x} <\,\pi^+(p_+)\pi^-(p_-)\pi^0(p_0)\,{\mbox out}\,\vert\, T\{j^{\mu}(x)j^{\nu}(0)\}\,\vert\,\Omega\,>\ .\lbl{red} \end{eqnarray} In the above expressions, $j_{\mu}(x)$ denotes the electromagnetic current, with its usual decomposition into an isotriplet and an isosinglet component, $j_{\mu}=j_{\mu}^3+j_{\mu}^0$, while $\vert\,\Omega\,>$ stands for the QCD vacuum with massive light quarks, but in the absence of electromagnetism (radiative corrections to the two processes \rf{proc} are not considered here). Since we also neglect isospin breaking effects due to $m_u\neq m_d$, Bose symmetry and G--parity constrain the three final pions to be in an $I=1$ total isospin state. Thus the matrix elements in Eq. \rf{red} only involve the $I=1$ component of the product of the two electromagnetic currents, \vs{2}\begin{equation} T\big\{j_{\mu}(x)j_{\nu}(0)\big\}_{I=1}\ = \ T\big\{j^3_{\mu}(x)j^0_{\nu}(0)+j^0_{\mu}(x)j^3_{\nu}(0)\big\}\ . \\[2mm]\end{equation} Therefore, isospin invariance relates the two amplitudes ${\cal A}^C$ and ${\cal A}^N$ in a simple way. With the Condon and Shortley phase convention adopted in Eq.~\rf{Ufield}, this relation reads \vs{2}\begin{equation} -{\cal A}^N(k,\epsilon;k',\epsilon';p_1,p_2,p_3)\ =\ {\cal A}^C(k,\epsilon;k',\epsilon';p_1,p_2,p_3)+ {\mbox{cyclic}}\,(p_1,p_2,p_3)\ , \lbl{iso} \\[2mm]\end{equation} where ``${\mbox{cyclic}}\,(p_1,p_2,p_3)$'' indicates that the contributions arising from cyclic permutations over the pion momenta $p_1, p_2$ and $p_3$ have to be added. Up to permutations of the momenta and polarizations of the photons, and/or permutations of the momenta of the charged pions, the amplitude ${\cal A}^C$ may be decomposed into six independent Lorentz invariant amplitudes \begin{eqnarray} &&{\cal A}^C(k,\epsilon;k',\epsilon';p_0,p_+,p_-) \ = \ {\cal A}^C_1(k,k';p_0,p_+,p_-)\, t_1(k,\epsilon;k',\epsilon')\nonumber\\ &&\quad +\ \sum_{i=2,3}\,\bigg[{\cal A}^C_i(k,k';p_0,p_+,p_-)\, t_i(k,\epsilon;k',\epsilon';p_0,p_+,p_-)\ + \ { k \choose \epsilon } \leftrightarrow { k' \choose \epsilon' } \bigg]\lbl{decomp}\\ &&\quad +\ \sum_{i=4,5,6}\,\bigg\{\bigg[{\cal A}^C_i(k,k';p_0,p_+,p_-)\, t_i(k,\epsilon;k',\epsilon';p_0,p_+,p_-)\ + \ { k \choose \epsilon } \leftrightarrow { k' \choose \epsilon' } \bigg] + \bigg[p_{+} \leftrightarrow p_{-} \bigg]\bigg\} \ ,\nonumber \end{eqnarray} with ($p_{ij}\equiv p_i+p_j$, $i,j=+,-,0$) \begin{eqnarray} & & t_1(k,\epsilon;k',\epsilon')= \epsilon_{\mu \nu \alpha \beta} k^\mu \epsilon^\nu k'^\alpha \epsilon'^\beta, \nonumber \\ & & t_2(k,\epsilon;k',\epsilon';p_0,p_+,p_-)=\epsilon_{\mu \nu \alpha \beta} k'^\mu \epsilon'^\nu p_{0}^\alpha \Bigl(-\epsilon^\beta+ {\epsilon \cdot p_{-}\over k\cdot p_{-}}p_{+}^{\beta} + {\epsilon \cdot p_{+}\over k\cdot p_{+}}p_{-}^{\beta}\Bigr), \nonumber \\ & & t_3(k,\epsilon;k',\epsilon';p_0,p_+,p_-)=\epsilon_{\mu \nu \alpha \beta} \Bigl(-\epsilon^\mu +{\epsilon \cdot p_{+-} \over k\cdot p_{+-}}k^\mu\Bigr) k'^\nu p_{0}^\alpha \epsilon'^\beta, \nonumber \\ & & t_4(k,\epsilon;k',\epsilon';p_0,p_+,p_-)=\epsilon_{\mu \nu \alpha \beta} \Bigl(-\epsilon^\mu +{\epsilon \cdot p_{+0} \over k\cdot p_{+0}}k^\mu\Bigr) k'^\nu p_{-}^\alpha \epsilon'^\beta, \nonumber \\ & & t_5(k,\epsilon;k',\epsilon';p_0,p_+,p_-)=\epsilon_{\mu \nu \alpha \beta} k'^\mu p_{+}^\nu p_{-}^\alpha \epsilon'^\beta \Bigl({\epsilon \cdot p_{+0} \over k \cdot p_{+0}} - {\epsilon \cdot p_{+} \over k \cdot p_{+}}\Bigr), \nonumber \\ & &t_6(k,\epsilon;k',\epsilon';p_0,p_+,p_-) =\epsilon_{\mu \nu \alpha \beta} \Bigl(-\epsilon^\mu +{\epsilon \cdot p_{+} \over k\cdot p_{+}}k^\mu \Bigr) k'^\nu p_{-}^\alpha \epsilon'^\beta\ . \end{eqnarray} In the sequel, we shall compute the amplitudes ${\cal A}^C$ and ${\cal A}^N$ within the framework of generalized chiral perturbation theory up to order one--loop. At this level of accuracy of the chiral expansion, the amplitude ${\cal A}^N$ does not yet receive its full structure as implied by Eqs. \rf{decomp} and \rf{iso}. Rather, it takes the simpler form ($p_{ij}\equiv p_i+p_j$, $i,j=1,2,3$) \begin{eqnarray} &&{\cal A}^N(k,\epsilon;k'\epsilon';p_1,p_2,p_3) \ = \ {\cal A}^N_1(k,k';p_1,p_2,p_3)\, t_1(k,\epsilon;k',\epsilon') \nonumber\\ &&\quad +\ \bigg[{\cal A}^N_2(k,k';p_1,p_2,p_3) \epsilon_{\mu \nu \alpha \beta} \Bigl(\epsilon'^\mu -{\epsilon' \cdot p_{12} \over k'\cdot p_{12}}k'^\mu\Bigr) p_{12}^{\nu}\epsilon^{\alpha}k^{\beta}\,+\,{\mbox{cyclic}}\,(p_1,p_2,p_3)\bigg] \nonumber\\ &&\quad + \ \bigg[{ k \choose \epsilon } \leftrightarrow { k' \choose \epsilon' } \bigg]\ +\ \cdots\ ,\lbl{Aneu} \end{eqnarray} where the ellipsis stands for higher order terms in the chiral expansion. In order that \rf{Aneu} follows from \rf{iso} and \rf{decomp}, it is {\it sufficient} that the one--loop charged amplitudes ${\cal A}_i^C$ satisfy the following conditions~: \begin{description} \item[i)] ${\cal A}_2^C(k,k';p_0,p_+,p_-)-{\cal A}_5^C(k,k';p_+,p_-,p_0)$ is entirely symmetric under permutations of the pion momenta $p_+, p_-, p_0$; \item[ii)] $k\cdot p_0\,\big[{\cal A}_5^C(k,k';p_0,p_+,p_-)+ {\cal A}_6^C(k,k';p_0,p_+,p_-)\big]\ =$ $\qquad\ k\cdot p_+ \,\big[{\cal A}_5^C(k,k';p_+,p_0,p_-)+ {\cal A}_6^C(k,k';p_+,p_0,p_-)\big]$; \item[iii)] ${\cal A}_5^C(k,k';p_0,p_+,p_-)\ =\ {\cal A}_5^C(k,k';p_+,p_0,p_-)$. \end{description} Furthermore, the two neutral amplitudes have then the following expressions in terms of the charged ones~: \begin{eqnarray} {\cal A}^N_1(k,k';p_1,p_2,p_3) &=& -{\cal A}^C_1(k,k';p_1,p_2,p_3)\nonumber\\ &&+\ \bigg[\,{1\over 3}{\cal A}^C_2(k,k';p_1,p_2,p_3)+ {\cal A}^C_3(k,k';p_3,p_1,p_2)\nonumber\\ &&+{\cal A}^C_4(k,k';p_1,p_2,p_3)+ {\cal A}^C_4(k,k';p_2,p_1,p_3)\nonumber\\ &&+{\cal A}^C_6(k,k';p_1,p_2,p_3)+ {\cal A}^C_6(k,k';p_2,p_1,p_3)\nonumber\\ &&+{2\over 3}{\cal A}^C_5(k,k';p_2,p_3,p_1)\,\bigg]\ + \ \bigg[{ k \choose \epsilon } \leftrightarrow { k' \choose \epsilon'} \bigg]\nonumber\\ &&+\ {\mbox{cyclic}}\,(p_1,p_2,p_3)\ ,\lbl{iso1} \end{eqnarray} and \begin{eqnarray} -{\cal A}^N_2(k,k';p_1,p_2,p_3) &=& \big[ {\cal A}^C_4(k,k';p_1,p_2,p_3)+{\cal A}^C_6(k,k';p_1,p_2,p_3)\big]+ \big[\,p_1\leftrightarrow p_2\,\big]\nonumber\\ &&+ {\cal A}^C_3(k,k';p_1,p_2,p_3)+{\cal A}^C_5(k,k';p_2,p_1,p_3)\ .\lbl{iso2} \end{eqnarray} \section{The Effective Lagrangian in GChPT} \setcounter{equation}{0} Before proceeding with the calculation of the amplitudes ${\cal A}^C$ and ${\cal A}^N$ in the next section, we first discuss the structure of the low--energy generating functional in the case of two light flavours that we need for our subsequent calculation. Since most of the results of this section are not available from the existing literature, we discuss them in some detail. The structure of the effective lagrangian ${\cal L}^{\mbox{\scriptsize eff}}$ is independent of the underlying mechanism of spontaneous chiral symmetry breaking. It consists of an infinite tower of chiral invariant contributions \vs{2}\begin{equation} {\cal L}^{\mbox{\scriptsize eff}} = \sum_{(k,l)}\,{\cal L}_{(k,l)}\ ,\lbl{Leff} \\[2mm]\end{equation} where ${\cal L}_{(k,l)}$ contains $k$ powers of covariant derivatives and $l$ powers of the scalar or pseudoscalar sources. In the chiral limit, these terms behave as \vs{2}\begin{equation} {\cal L}_{(k,l)}\sim\left({{\mbox p}\over{\Lambda_H}}\right)^k \left({{m_{\mbox{\scriptsize quark}}}\over{\Lambda_H}}\right)^l\ , \\[2mm]\end{equation} with $m_{\mbox{\scriptsize quark}}=m_u,m_d$, and p stands for a typical external momentum. The standard approach not only assumes $m_{\mbox{\scriptsize quark}}\ll\Lambda_H$, but also $m_{\mbox{\scriptsize quark}}\ll B/2A$, such as to enforce the dominance of the term linear in $m_{\mbox{\scriptsize quark}}$ in the expression of the pion mass~\rf{Mpilead}. This allows to reorganize the double expansion \rf{Leff} as~\cite{GL1,GL2} \vs{2}\begin{equation} {\cal L}^{\mbox{\scriptsize eff}} = {\cal L}^{(2)}+{\cal L}^{(4)}+{\cal L}^{(6)}+\cdots\ , \\[2mm]\end{equation} where ${\cal L}^{(d)}=\sum\,{\cal L}_{(k,l)}$ with $k+2l=d$. The generalized framework considers the possibility that the condensate could be much smaller than usually believed, so that for the actual values of the quark masses one could have $m_{\mbox{\scriptsize quark}}\sim B/2A$ and still $m_{\mbox{\scriptsize quark}}\ll\Lambda_H$. This leads to a different reorganization of the double expansion \rf{Leff}, namely \vs{2}\begin{equation} {\cal L}^{\mbox{\scriptsize eff}} = {\tilde{\cal L}}^{(2)}+{\tilde{\cal L}}^{(3)}+{\tilde{\cal L}}^{(4)} +{\tilde{\cal L}}^{(5)}+{\tilde{\cal L}}^{(6)}+\cdots\ , \\[2mm]\end{equation} where now, according to \rf{gencount}, ${\tilde{\cal L}}^{(d)}=\sum\,B^n{\cal L}_{(k,l)}$ with $k+l+n=d$ \cite{pionscat,gchpt}. The leading order of the generalized expansion is described by ${\tilde{\cal L}}^{(2)}$, which in the two flavour case was given in Section 1, Eq.~\rf{L2}. For our purposes, we need only to consider the situation without axial source, and with the vector source restricted to the (classical) photon field $A_{\mu}$, \vs{2}\begin{equation} a_{\mu}\ =\ 0\ ,\ \ v_{\mu}\ =\ eA_{\mu}Q\ , \\[2mm]\end{equation} where $Q$ stands for the charge matrix of the two light quark flavours $u$ and $d$, \vs{2}\begin{equation} Q\ =\ \left( \begin{array}{cc} {2\over 3} & 0\\ 0 & -{1\over 3}\end{array}\right)\ . \\[2mm]\end{equation} In GChPT, the next--to--leading--order corrections are of order ${\cal O}({\mbox p}^3)$, and still occur before the loop corrections. They are embodied in ${\tilde{\cal L}}^{(3)}={\cal L}_{(2,1)}+{\cal L}_{(0,3)}$, which reads~\footnote{~The superscript (2) is meant to distinguish the low energy constants $\xi^{(2)}$, $\rho_i^{(2)}$ from the similar ones that occur in the expression of ${\tilde{\cal L}}^{(3)}$ in the three flavour case \cite{pionscat,gchpt}.} \begin{eqnarray} \tilde{\cal L}^{(3)}&=&{1\over 4}F^2 \left\{\xi^{(2)}\langle D_\mu U^+D^\mu U(\chi^+U+U^+\chi)\rangle \right.\nonumber\\ &&\qquad + \rho_1^{(2)}\langle (\chi^+U)^3+(U^+\chi)^3\rangle +\rho_2^{(2)}\langle (\chi^+U+U^+\chi)\chi^+\chi\rangle\nonumber\\ &&\qquad + \rho_3^{(2)}\langle\chi^+U-U^+\chi\rangle \langle(\chi^+U)^2-(U^+\chi)^2\rangle\lbl{L3} \\ &&\qquad + \rho_4^{(2)}\langle(\chi^+U)^2 +(U^+\chi)^2\rangle \langle\chi^+U+U^+\chi\rangle\nonumber\\ &&\qquad \left. + \rho_5^{(2)}\langle\chi^+\chi\rangle \langle\chi^+U+U^+\chi\rangle \right\}\ .\nonumber \end{eqnarray} >From order ${\cal O}({\mbox p}^4)$ onward, the contributions to ${\cal L}^{\mbox{\scriptsize eff}}$ come with either even or odd intrinsic parity, ${\tilde{\cal L}}^{(d)} = {\tilde{\cal L}}^{(d)}_+ +{\tilde{\cal L}}^{(d)}_-$ for $d\ge 4$, or ${\cal L}_{(k,l)}={\cal L}_{(k,l)}^{+}+{\cal L}_{(k,l)}^{-}$ for $k\ge 4$. The tree--level contributions at order ${\cal O}({\mbox p}^4)$ in the {\it even intrinsic parity sector} are contained in \vs{2}\begin{equation} \tilde{\cal L}^{(4)}_+={\cal L}_{(4,0)}^{+}+{\cal L}_{(2,2)}+{\cal L}_{(0,4)}+ B^2{\cal L}'_{(0,2)} + B{\cal L}'_{(2,1)} + B{\cal L}'_{(0,3)}\ . \lbl{L4} \\[2mm]\end{equation} The part without explicit chiral symmetry breaking, ${\cal L}_{(4,0)}^{+}$, is described by the same low--energy constants $l_1$, $l_2$, $l_5$, $l_6$ and $h_2$ as in the standard case \cite{GL1}, \begin{eqnarray} {\cal L}_{(4,0)}^{+} &=& {{l_1}\over 4}\,\langle\,D^{\mu}U^+D_{\mu}U\, \rangle^2\,+\,{{l_2}\over 4}\,\langle\,D^{\mu}U^+D^{\nu}U\, \rangle\,\langle\,D_{\mu}U^+D_{\nu}U\,\rangle \nonumber\\ && +\,l_5\,\langle \,F^{\mbox{\scriptsize L}}_{\mu\nu}UF^{{\mbox{\scriptsize L}}\,\mu\nu}U^+ \,\rangle \,+\,{{il_6}\over 2}\,\langle \,F^{\mbox{\scriptsize R}}_{\mu\nu}d^{\mu}Ud^{\nu}U^+ + F^{\mbox{\scriptsize L}}_{\mu\nu}d^{\mu}U^+d^{\nu}U\,\rangle \nonumber\\ && -\,(2h_2+\frac{1}{2}l_5)\,\langle \,F^{\mbox{\scriptsize R}}_{\mu\nu}F^{{\mbox{\scriptsize R}}\,\mu\nu} +F^{\mbox{\scriptsize L}}_{\mu\nu}F^{{\mbox{\scriptsize L}}\,\mu\nu} \,\rangle\ . \end{eqnarray} The part with two powers of momenta and two powers of quark masses is given by \begin{eqnarray} {\cal L}_{(2,2)}&=& {1\over{4}}F^2\bigg\{ a_1 \langle D_{\mu}U^+D^{\mu}U (\chi^+\chi + U^+\chi\chi^+U)\rangle \nonumber\\ & &\qquad + a_2 \langle D_{\mu}U^+U\chi^+D^{\mu}UU^+\chi\rangle \nonumber\\ & &\qquad + a_3 \langle D_{\mu}U^+U(\chi^+D^{\mu}\chi-D^{\mu}\chi^+\chi) + D_{\mu}UU^+(\chi D^{\mu}\chi^+ - D^{\mu}\chi\chi^+)\rangle \nonumber\\ & &\qquad + b_1 \langle D_{\mu}U^+D^{\mu}U (\chi^+U\chi^+U + U^+\chi U^+\chi)\rangle \nonumber\\ & &\qquad + b_2 \langle D_{\mu}U^+\chi D^{\mu}U^+\chi + \chi^+D_{\mu}U\chi^+D^{\mu}U\rangle\nonumber\\ & &\qquad + b_3 \langle U^+D_{\mu}\chi U^+D^{\mu}\chi+D_{\mu}\chi^+UD^{\mu}\chi^+U\rangle\nonumber\\ & &\qquad + c_1 \langle D_{\mu}U^+\chi +\chi^+ D_{\mu}U\rangle \langle D^{\mu}U^+\chi +\chi^+ D^{\mu}U\rangle\nonumber\\ & &\qquad + c_2 \langle D_{\mu}\chi^+U + U^+ D_{\mu}\chi\rangle \langle D^{\mu}U^+\chi +\chi^+ D^{\mu}U\rangle\nonumber\\ & &\qquad + c_3 \langle D_{\mu}\chi^+U + U^+ D_{\mu}\chi\rangle \langle D^{\mu}\chi^+U + U^+D^{\mu}\chi\rangle \lbl{L22}\\ & &\qquad + c_4 \langle D_{\mu}U^+\chi -\chi^+ D_{\mu}U\rangle \langle D^{\mu}U^+\chi -\chi^+ D^{\mu}U\rangle\nonumber\\ & &\qquad + c_5 \langle D_{\mu}\chi^+U - U^+D_{\mu}\chi\rangle \langle D^{\mu}\chi^+U - U^+D^{\mu}\chi\rangle\nonumber\\ & &\qquad + h_3 \langle D_{\mu}\chi^+D^{\mu}\chi^+\rangle \bigg\} \ .\nonumber \end{eqnarray} Finally, the tree--level contributions which behave as ${\cal O} (m_{\mbox{\scriptsize quark}}^4)$ in the chiral limit are contained in ${\cal L}_{(0,4)}$, which reads \begin{eqnarray} {\cal L}_{(0,4)} &=& {1\over{4}}F^2\bigg\{ e_1 \langle (\chi^+U)^4 + (U^+\chi)^4 \rangle \nonumber\\ & &\qquad + e_2 \langle \chi^+\chi(\chi^+U\chi^+U+U^+\chi U^+\chi) \rangle\nonumber\\ & &\qquad + e_3 \langle \chi^+\chi U^+\chi\chi^+ U \rangle \nonumber\\ & &\qquad + f_1 \langle (\chi^+U)^2 + (U^+\chi)^2 \rangle^2\nonumber\\ & &\qquad + f_2 \langle (\chi^+U)^3 + (U^+\chi)^3 \rangle\langle \chi^+U + U^+\chi\rangle \nonumber\\ & &\qquad + f_3 \langle \chi^+\chi(\chi^+U + U^+\chi)\rangle \langle\chi^+U + U^+\chi\rangle\nonumber\\ & &\qquad + f_4 \langle (\chi^+U)^2 + (U^+\chi)^2 \rangle \langle \chi^+U + U^+\chi\rangle^2 \nonumber\\ & &\qquad + f_5 \langle (\chi^+U)^3 - (U^+\chi)^3 \rangle\langle \chi^+U - U^+\chi\rangle \nonumber\\ & &\qquad + h_4 \langle\chi^+\chi\chi^+\chi \rangle\nonumber\\ & &\qquad + h_5 \langle\chi^+\chi \rangle ({\mbox{det}}\chi + {\mbox{det}}\chi^+ )\nonumber\\ & &\qquad + h_6 ({\mbox{det}}\chi + {\mbox{det}}\chi^+ )^2 \nonumber\\ & &\qquad + h_7 ({\mbox{det}}\chi - {\mbox{det}}\chi^+ )^2 \bigg\} \ . \nonumber \end{eqnarray} Notice that in the standard framework the contributions from ${\cal L}_{(2,2)}$ and from ${\cal L}_{(0,4)}$would count as order ${\cal O}({\mbox p}^6)$ and order ${\cal O}({\mbox p}^8)$, respectively~\footnote{~The standard ${\cal O}({\mbox p}^6)$ effective lagrangian ${\cal L}^{(6)}={\cal L}_{(6,0)}+{\cal L}_{(4,1)}+{\cal L}_{(2,2)}+{\cal L}_{(0,3)}$ has been worked out in Ref. \cite{FS} for the case of three light flavours, and very recently, for both two and three light flavours, in Ref.~\cite{bijnensetal}.}. Next, we turn to the {\it odd intrinsic parity sector}. There, the first contribution starts at order ${\cal O}({\mbox p}^4)$, and since it is entirely fixed by the short distance properties of QCD, there is no difference between the standard and the generalized case, ${\cal L}^{(4)}_-={\tilde{\cal L}}^{(4)}_-= {\cal L}_{(4,0)}^-$. In the two flavour case, ${\cal L}_{(4,0)}^-$ vanishes in the absence of external sources. In the presence of an electromagnetic field, it reads \begin{eqnarray} {\cal L}_{(4,0)}^-&=& {e\over{16\pi^2}}\epsilon^{\mu\nu\alpha\beta}A_{\mu}\langle Q\big(\partial_\nu U\partial_\alpha U^+\partial_\beta UU^+- \partial_\nu U^+\partial_\alpha U\partial_\beta U^+U\big)\rangle \\ && -{{ie^2}\over{8\pi^2}}\epsilon^{\mu\nu\alpha\beta}\partial_\mu A_\nu A_\alpha \langle Q^2\partial_\beta UU^++Q^2U^+\partial_\beta U -{1\over 2}QUQ\partial_\beta U^+ +{1\over 2}QU^+Q\partial_\beta U\rangle\ . \nonumber \end{eqnarray} The computation of the amplitudes ${\cal A}^N$ and ${\cal A}^C$ to one loop also involves the counterterms from ${\tilde{\cal L}}^{(5)}_-={\cal L}_{(4,1)}^-$ and from ${\tilde{\cal L}}^{(6)}_-={\cal L}_{(6,0)}^-+{\cal L}_{(4,2)}^-$. In the standard case, both ${\cal L}_{(4,1)}^-$ and ${\cal L}_{(6,0)}^-$ count as order ${\cal O}({\mbox p}^6)$, and have been discussed before in the literature \footnote{For a review, see \cite{bijnensrev}.} \cite{Issl,AkAl,FS}. Borrowing from the last and most recent of these references, we obtain \begin{eqnarray} {\cal L}_{(4,1)}^- &=& {1\over{4\pi^2}}\epsilon_{\mu\nu\alpha\beta}\bigg\{ iA_{4} \langle\cm{\chi}\cp{G^{\mu\nu}}\cp{G^{\alpha\beta}}\rangle + iA_{6} \langle\cm{\chi}\rangle\langle\cp{G^{\mu\nu}}\cp{G^{\alpha\beta}}\rangle \nonumber\\ && + A_{12} \langle\cm{D^\mu U}\cm{D^\nu U}(\cm{\chi}\cp{G^{\alpha\beta}} + \cp{G^{\alpha\beta}}\cm{\chi})\rangle \nonumber\\ && + A_{13} \langle\cm{D^\mu U}\cm{\chi}\cm{D^\nu U}\cp{G^{\alpha\beta}}\rangle \,\cdots\bigg\}\ ,\lbl{L41-} \end{eqnarray} and \begin{eqnarray} {\cal L}_{(6,0)}^- &=& {1\over{4\pi^2}}\epsilon_{\mu\nu\alpha\beta}\bigg\{ iA_{2} \langle\cm{D^\mu U}(\cp{D^\nu G^{\gamma\alpha}}\cp{{G_\gamma}^\beta} - \cp{G^{\gamma\alpha}}\cp{D^\nu {G_\gamma}^\beta})\rangle \nonumber\\ && + iA_{3} \langle\cm{D^\mu U}(\cp{D_\gamma G^{\gamma\nu}}\cp{G^{\alpha\beta}} - \cp{G^{\gamma\nu}}\cp{D_\gamma G^{\alpha\beta}} \nonumber\\ &&\qquad\qquad\qquad\qquad -\cp{D_\gamma G^{\alpha\beta}}\cp{G^{\gamma\nu}} + \cp{G^{\alpha\beta}}\cp{D_\gamma G^{\gamma\nu}}) \rangle \nonumber\\ && + A_{7} \langle\cm{D^\alpha D^\gamma U}(\cm{D_\gamma U}\cm{D^\beta U}\cp{G^{\mu\nu}} -\cp{G^{\mu\nu}}\cm{D^\beta U}\cm{D_\gamma U})\rangle \nonumber\\ && + A_{8} \langle\cm{D^\alpha D^\gamma U}(\cm{D^\beta U}\cm{D_\gamma U}\cp{G^{\mu\nu}} - \cp{G^{\mu\nu}}\cm{D_\gamma U}\cm{D^\beta U})\rangle \nonumber\\ && +\ \cdots\bigg\}\ . \lbl{L60-} \end{eqnarray} \noindent Here, we have only listed those terms that will actually contribute to the processes under study, when the mass--shell conditions for the momenta and polarizations of the photons are taken into account. We have however kept the numbering of the low-energy constants introduced in \cite{FS}, but we have, for convenience, changed their normalization by an overall factor $1/4\pi^2$. The notation is otherwise as in \cite{FS}, except for the fact that the source $\chi$ does not contain the factor $2B$, see Eq. \rf{chi}. The last piece we need for a full one--loop computation of the amplitudes \rf{red} is ${\cal L}_{(4,2)}^-$. It counts as order ${\cal O}({\mbox p}^8)$ in the {\it standard} case, and is not available from the existing literature. These contributions, which are order ${\cal O}({\widehat m}^2)$ corrections to ${\cal L}_{(4,0)}^-$, are expected to be small in the two--flavour chiral expansion, and will be parametrized appropriately in the one--loop expressions of the amplitudes ${\cal A}^C$ and ${\cal A}^N$ given in the next section. The determination of the combinations of low--energy constants that enter these amplitudes will be discussed in Section 5 below. When studying a given process one also needs to take into account contributions from pion loops, which produce divergences that are eliminated by a renormalization of the low--energy constants of the effective lagrangian. We have computed this divergent part of the one--loop generating functional in the even intrinsic parity sector using standard heat--kernel techniques, and we have then performed the corresponding renormalization of the low--energy constants in the same dimensional renormalization scheme as described in \cite{GL1,GL2}. Thus the low--energy constants display a logarithmic scale dependence ($X(\mu )$ denotes generically any of these renormalized low-energy constants) \vs{2}\begin{equation} X(\mu )= X(\mu ') + {{\Gamma_X}\over{(4\pi )^2}}\,\cdot\ln({\mu '}/\mu )\ . \lbl{scale} \\[2mm]\end{equation} The full list of the resulting $\beta$--function coefficients $\Gamma_X$ is given in Table 1 below.\footnote{~These results have also been established independently by L.~Girlanda, private communication to M.K. and \cite{girl98}. The renormalization of the ${\cal L}_{(4,0)}^+$ counterterms has, of course, already been obtained before in \cite{GL1}.} \begin{center} \begin{tabular}{c|ccc|ccc|c} $X$ & $F^2\cdot\Gamma_X$&$\qquad\qquad$&$X$ & $F^2\cdot\Gamma_X$& $\qquad\qquad$&$X$ & $\Gamma_X$\\ \cline{1-2} \cline{4-5}\cline{7-8} $A$ & $3B^2$ & & $\xi^{(2)}$ & $4B$& &$l_1$ & ${1\over 3}$\\ \cline{4-5} $Z^P$ & $-{3\over 2}B^2$ & &$\rho_1^{(2)}$ & $-4B(A+Z^P)$ & &$l_2$ &${2\over 3}$ \\ \cline{1-2} $h_0$ & $0$ & &$\rho_2^{(2)}$ & $-4B(A-3Z^P)$ & &$l_5$ & $-{1\over 6}$\\ $h_1$ & $6B^2$ & &$\rho_3^{(2)}$ & $2B(A+3Z^P)$ & &$l_6$ & $-{1\over 3}$\\ \cline{7-8} & & &$\rho_4^{(2)}$ & $2B(3A+Z^P)$& &$h_2$ & ${1\over 12}$\\ & & &$\rho_5^{(2)}$ & $4B(A-2Z^P)$& & \end{tabular} \vskip 0.5 true cm {\bf Table 1a:} {\it Scale dependence of the low energy constants of ${\tilde{\cal L}}^{(2)}$, ${\tilde{\cal L}}^{(3)}$ and of ${{\cal L}}_{(4,0)}^{+}$.} \end{center} \begin{center} \begin{tabular}{c|ccc|c} $X$ & $F^2\cdot\Gamma_X$ & & $X$ & $F^2\cdot\Gamma_X$\\ \cline{1-2}\cline{4-5} $a_1$ & $-2Z^P$ &$\qquad$ &$e_1$ & $-4A^2-22(Z^P)^2-20AZ^P$\\ $a_2$ & $-12Z^P$ &$\qquad$ &$e_2$ & $-4A^2-16(Z^P)^2-12AZ^P$\\ $a_3$ & $0$ &$\qquad$ &$e_3$ & $-12A^2-64(Z^P)^2-48AZ^P$\\ \cline{1-2}\cline{4-5} $b_1$ & $6(A+Z^P)$ &$\qquad$ &$f_1$ & $3A^2+15(Z^P)^2+12AZ^P$\\ $b_2$ & $-2(A+Z^P)$&$\qquad$ &$f_2$ & $2A^2+8(Z^P)^2+10AZ^P$\\ $b_3$ & $0$ &$\qquad$ &$f_3$ & $4A^2+24(Z^P)^2+20AZ^P$\\ \cline{1-2} $c_1$ & $2(A+2Z^P)$&$\qquad$ &$f_4$ & $-6(Z^P+A)Z^P$\\ $c_2$ & $0$ &$\qquad$ &$f_5$ & $2A^2+8(Z^P)^2+10AZ^P$\\ \cline{4-5} $c_3$ & $0$ &$\qquad$ &$h_4$ & $4A^2+8(Z^P)^2+8AZ^P$\\ $c_4$ & $2A$ &$\qquad$ &$h_5$ & $-4A^2-32(Z^P)^2-28AZ^P$\\ $c_5$ & $0$ &$\qquad$ &$h_6$ & $4A^2+6(Z^P)^2+8AZ^P$\\ \cline{1-2} $h_3$ & $0$ &$\qquad\qquad$ &$h_7$ & $-14(Z^P)^2$\\ \end{tabular} \vskip 0.5 true cm {\bf Table 1b:} {\it Scale dependence of the low energy constants of ${\cal L}_{(2,2)}$ and of ${\cal L}_{(0,4)}$.} \end{center} Notice that at order ${\cal O}({\mbox p}^4)$, the low-energy constants of ${\tilde{\cal L}}^{(2)}$ and ${\tilde{\cal L}}^{(3)}$ also need to be renormalized. The corresponding counterterms, however, are of order ${\cal O}(B^2)$ and ${\cal O}(B)$, respectively, and they are gathered in the three last terms of Eq. \rf{L4}~: in GChPT, renormalisation proceeds order by order in the expansion in powers of $B$. Alternatively, one may think of Eqs. \rf{L2} and \rf{L3} as standing for the combinations ${\tilde{\cal L}}^{(2)} + B^2{\cal L}'_{(0,2)}$ and ${\tilde{\cal L}}^{(3)} + B{\cal L}'_{(2,1)} + B{\cal L}'_{(0,3)}$, respectively, with the corresponding low-energy constants representing the renormalized, scale dependent, quantities. We shall adopt the latter point of view in the sequel. We have not worked out the general structure of the divergent part of the one--loop generating functional in the odd intrinsic parity sector. For SChPT, this has been done in Refs.~\cite{DonWyl,Issl,BijBraCor}. \section{The One--Loop Amplitudes} \setcounter{equation}{0} Having constructed the effective lagrangian in the preceding section, the computation of the amplitudes ${\cal A}^N$ and ${\cal A}^C$ at next--to--leading order is a straightforward exercise. We begin with the one--loop expression of the neutral amplitude ${\cal A}^N$, whose structure at that level of the chiral expansion is given by \rf{Aneu}, with \begin{eqnarray} {\cal A}^N_1(k,k';p_1,p_2,p_3)&=&{{{\cal A}^{\pi^0\to\gamma\gamma} {\cal A}^{00;00}(p_{12}^2,p_{13}^2,p_{23}^2)}\over{s-M_\pi^2}} \nonumber\\ && -\frac{e^2}{4 \pi^2 F_\pi^3}\bigg\{ \frac{1}{2}(\beta -1)+\frac{1}{2}{\widehat m} t (\beta -3) -\alpha M_{\pi}^2t'-4{\widehat m}^2t''-\gamma_{00} \nonumber\\ && +\frac{1}{F_\pi^2} (\lambda_1+2\lambda_2)(s-3M_\pi^2)) \nonumber\\ && +\frac{1}{F_\pi^2} {\bar J}(p_{12}^2)\big[p_{12}^2-M_\pi^2+ {1\over 3}(\alpha-1)M_\pi^2\big] \nonumber\\ && +\frac{1}{F_\pi^2} {\bar J}(p_{13}^2)\big[p_{13}^2-M_\pi^2+ {1\over 3}(\alpha-1)M_\pi^2\big] \nonumber\\ && +\frac{1}{F_\pi^2} {\bar J}(p_{23}^2)\big[p_{23}^2-M_\pi^2+ {1\over 3}(\alpha-1)M_\pi^2\big] \nonumber\\ && +\frac{2}{F_\pi^2} \big[{\bar R}(p_{12}^2,k\cdot p_{12})+{\bar R}(p_{12}^2,k'\cdot p_{12})\big] \big[p_{12}^2-M_\pi^2+\frac{1}{3}(\alpha-1)M_\pi^2\big] \nonumber\\ && +\frac{2}{F_\pi^2} \big[{\bar R}(p_{13}^2,k\cdot p_{13})+{\bar R}(p_{13}^2,k'\cdot p_{13})\big] \big[p_{13}^2-M_\pi^2+\frac{1}{3}(\alpha-1)M_\pi^2\big] \nonumber\\ && +\frac{2}{F_\pi^2} \big[{\bar R}(p_{23}^2,k\cdot p_{23})+{\bar R}(p_{23}^2,k'\cdot p_{23})\big] \big[p_{23}^2-M_\pi^2+\frac{1}{3}(\alpha-1)M_\pi^2\big] \bigg\}\ , \nonumber\\ && \lbl{A1N} \end{eqnarray} and \vs{2}\begin{equation} {\cal A}^N_2(k,k';p_1,p_2,p_3)= \frac{e^2}{2\pi^2 F_\pi^5}{\bar R}(p_{12}^2,k\cdot p_{12}) \big[p_{12}^2-M_\pi^2+{1\over 3}(\alpha -1)M_\pi^2\big]\ .\lbl{A2N} \\[2mm]\end{equation} This second amplitude, which was absent at tree level, is entirely generated by the pion loops. The numerator of the contribution which develops a pole at $s=M_\pi^2$ in the expression \rf{A1N} of ${\cal A}^N_1$ is given by the product of ${\cal A}^{\pi^0\to\gamma\gamma}$, which is related to the on-shell amplitude of the $\pi^0\to\gamma\gamma$ decay through ${\cal A}(\pi^0\to\gamma\gamma)=-t_1(k,\epsilon,k',\epsilon '){\cal A}^{\pi^0\to\gamma\gamma}$, times the amplitude ${\cal A}^{00;00}(p_{12}^2,p_{13}^2,p_{23}^2)$ of virtual $\pi^0-\pi^0$ scattering ($p_{12}^2+p_{13}^2+p_{23}^2=s+3M_\pi^2$). The expressions, at order ${\cal O}({\mbox p}^6)$ and order ${\cal O}({\mbox p}^4)$, respectively, of these amplitudes read \vs{2}\begin{equation} {\cal A}^{\pi^0\to\gamma\gamma}=\frac{e^2}{4 \pi^2 F_\pi} \big[1+{\widehat m} t+ M_\pi^2t' +{\widehat m}^2 t'' +2{\widehat m}^2a_3\big]\ ,\lbl{ggpi0} \\[2mm]\end{equation} and \begin{eqnarray} {\cal A}^{00;00}(p_{12}^2,p_{13}^2,p_{23}^2) &=& {{\alpha M_\pi^2}\over{F_\pi^2}} +\frac{1}{F_\pi^4} (\lambda_1+2\lambda_2)\big[(p_{12}^2-2M_\pi^2)^2 +(p_{13}^2-2M_\pi^2)^2+(p_{23}^2-2M_\pi^2)^2\big] \nonumber\\ &&+\frac{1}{F_\pi^4} {\bar J}(p_{12}^2)\bigg[(p_{12}^2-\frac{4}{3}M_\pi^2 +\frac{\alpha}{3}M_\pi^2)^2+\frac{\alpha^2}{2}M_\pi^4\bigg] \nonumber\\ && +\frac{1}{F_\pi^4} {\bar J}(p_{13}^2)\bigg[(p_{13}^2-\frac{4}{3}M_\pi^2 +\frac{\alpha}{3}M_\pi^2)^2+\frac{\alpha^2}{2}M_\pi^4\bigg] \nonumber\\ && +\frac{1}{F_\pi^4} {\bar J}(p_{23}^2)\bigg[(p_{23}^2-\frac{4}{3}M_\pi^2 +\frac{\alpha}{3}M_\pi^2)^2+\frac{\alpha^2}{2}M_\pi^4\bigg]\ .\lbl{A0000} \end{eqnarray} The various parameters $\alpha$, $\beta$, $\lambda_{1,2}$ and $\gamma_{00}$ involved in the expressions \rf{A1N} and \rf{A0000} are given in terms of combinations of the low--energy constants of the effective lagrangian ${\cal L}^{\mbox{\scriptsize eff}}$ and of chiral logarithms due to the pion loops. They read \begin{eqnarray} {{F_\pi^2}\over{F^2}}M_\pi^2\alpha &=& 2{\widehat m} B + 16{\widehat m}^2A \nonumber\\ && +{\widehat m}^3(81\rho_1^{(2)}+\rho_2^{(2)}+164\rho_4^{(2)}+2\rho_5^{(2)}) -4M_\pi^2{\widehat m}\xi^{(2)} \nonumber\\ && +16{\widehat m}^4(16e_1+e_2+32f_1+34f_2+2f_3+72f_4+6a_3A) \nonumber\\ && -8M_\pi^2{\widehat m}^2(2b_1-2b_2-a_3-4c_1) \nonumber\\ && -{1\over{16\pi^2F_\pi^2}} \big[ 2M_\pi^4 + 102{\widehat m}^2M_\pi^2A+264{\widehat m}^4A^2\big] \ln{{M_\pi^2}\over{\mu^2}} \nonumber\\ && -{1\over{16\pi^2F_\pi^2}} \big[ {{M_\pi^4}\over 2} + 44{\widehat m}^2M_\pi^2A+264{\widehat m}^4A^2\big]\ , \lbl{alphaexp} \end{eqnarray} \begin{eqnarray} \beta &=& 1 + 2{\widehat m}\xi^{(2)} - 4{\widehat m}^2(\xi^{(2)})^2 +2{\widehat m}^2(3a_2+2a_3+4b_1+2b_2+4c_1) \nonumber\\ && -{{M_\pi^2}\over{48\pi^2F_\pi^2}}\bigg(\ln{{M_\pi^2}\over{\mu^2}} +1\bigg) \big[ 6+5(\alpha -1)\big]\ , \nonumber\\ && \end{eqnarray} \begin{eqnarray} \lambda_1 &=& {1\over{48\pi^2}}\big( {\bar l}_1 - {4\over 3} \big)\ , \nonumber\\ &&\lbl{l1l2}\\ \lambda_2 &=& {1\over{48\pi^2}}\big( {\bar l}_2 - {5\over 6} \big)\ , \nonumber \end{eqnarray} \begin{eqnarray} \gamma_{00}&=&{\widehat m}^2(3a_2+2a_3+6b_2+12c_1) \nonumber\\ &&-{{M_\pi^2}\over{32\pi^2F_\pi^2}}\bigg(\ln{{M_\pi^2}\over{\mu^2}}+1\bigg) (\alpha-1)\ .\lbl{g00} \end{eqnarray} The parameters $t$, $t'$ and $t''$ contain the contributions from ${\cal L}_{(4,1)}^-$, ${\cal L}_{(6,0)}^-$ and ${\cal L}_{(4,2)}^-$, respectively. In particular, from the formulae \rf{L41-} and \rf{L60-} we derive \begin{eqnarray} t &=& {32\over 3}A_4\ , \nonumber\\ t' &=& -{8\over 3}(A_2-2A_3)\ .\lbl{ttprime} \end{eqnarray} Finally, ${\bar J}(s)$ denotes the Chew--Mandelstam function~\cite{chew}, the usual scalar two--point loop integral subtracted at $s=0$ (for its expression, see Ref.~\cite{GL1}), whereas ${\bar R}(p^2,k\cdot p)$, which is related to the three--point scalar loop function, is given, for $k^2=0$, by \vs{2}\begin{equation} {\bar R}(p^2,k\cdot p)={\bar C}(p^2, k\cdot p)- {{(k-p)^2}\over{4(k\cdot p)}}\big[{{\bar J}((k-p)^2)-\bar J}(p^2)\big] +{1\over{32\pi^2}}\ , \\[2mm]\end{equation} with ($\sigma = \sqrt{1-4M_\pi^2/p^2}$, $\sigma' = \sqrt{1-4M_\pi^2/(k-p)^2}$) \vs{2}\begin{equation} 16\pi^2{\bar C}(p^2, k\cdot p) = {{M_\pi^2}\over{4(k\cdot p)}}\big[ \ln^2\big({{\sigma-1}\over{\sigma+1}}\big) - \ln^2\big({{\sigma'-1}\over{\sigma'+1}}\big)\big]\ . \\[2mm]\end{equation} For the charged amplitude ${\cal A}^C$, the general structure is more involved, see Eq.~\rf{decomp}, and at one loop we obtain \begin{eqnarray} {\cal A}^C_1(k,k';p_0,p_+,p_-)&=& {{\cal A}^{\pi^0\to\gamma\gamma} {\cal A}^{00;+-}(p_{+-}^2,p_{+0}^2,p_{-0}^2)\over{s-M_\pi^2}} \nonumber\\ &&+\frac{e^2}{4\pi^2F_\pi^3}\bigg\{ \frac{1}{2}(\beta +1)+\frac{1}{2}{\widehat m} t(\beta -1) -\frac{1}{3}(\alpha -1)M_\pi^2t'\nonumber\\ &&+ \gamma_{+-}+\frac{2}{3}\gamma_{+-}'\nonumber\\ &&+ {{\lambda_1}\over{F_\pi^2}}(p_{+-}^2-2M_\pi^2) + {1\over{F_\pi^2}}(\lambda_2-{1\over{288\pi^2}}) (p_{+0}^2+p_{-0}^2-4M_\pi^2) \nonumber\\ && +{1\over{6F_\pi^2}} \bar{J}(p_{+-}^2)\big[3p_{+-}^2+4(\alpha -1)M_\pi^2\big] \nonumber\\ && +{1\over{12F_\pi^2}} \bar{J}(p_{+0}^2)\big[5p_{+0}^2-14M_\pi^2-2(\alpha -1)M_\pi^2\big] \nonumber\\ && +{1\over{12F_\pi^2}} \bar{J}(p_{-0}^2)\big[5p_{-0}^2-14M_\pi^2-2(\alpha -1)M_\pi^2\big]\bigg\}\ , \end{eqnarray} \begin{eqnarray} {\cal A}^C_2(k,k';p_0,p_+,p_-)&=& \frac{e^2}{24\pi^2F_\pi^3}\bigg\{ 6+6\gamma_{+-}'-\frac{1}{48\pi^2F_\pi^2}\big[(k'-p_0)^2+p_{+0}^2+p_{-0}^2\big] \nonumber\\ && -{1\over{F_\pi^2}}{\bar J}(p_{+0}^2)\big[4M_\pi^2-p_{+0}^2\big] -{1\over{F_\pi^2}}{\bar J}(p_{-0}^2)\big[4M_\pi^2-p_{-0}^2\big] \nonumber\\ && -{1\over{F_\pi^2}}{\bar J}\big((k'-p_0)^2\big)\big[4M_\pi^2-(k'-p_0)^2\big] \bigg\}\ , \end{eqnarray} \vs{2}\begin{equation} {\cal A}^C_3(k,k';p_0,p_+,p_-)= -\frac{e^2}{4\pi^2F_\pi^5} {\bar R}(p_{+-}^2;k\cdot p_{+-})\big[p_{+-}^2+{4\over 3}(\alpha -1)M_\pi^2 \big]\ , \\[2mm]\end{equation} \begin{eqnarray} {\cal A}^C_4(k,k';p_0,p_+,p_-)&=& -\frac{e^2}{4\pi^2F_\pi^5}\bigg\{ \frac{1}{3}\bar{J}\big((k'-p_{-})^2\big) M^2_\pi (4\frac{k\cdot p_+}{k\cdot p_{+0}}-1) \nonumber\\ && +\frac{1}{6} \big[\bar{J}\big((k'-p_{-})^2\big)-\bar{J}(p_{+0}^2)\big] \bigg[p_{+0}^2+2 M^2_\pi+\frac{1}{2}\frac{p_{+0}^2}{k\cdot p_{+0}} (4M^2_\pi-p_{+0}^2) \nonumber\\ &&\qquad +\frac{k\cdot p_+}{k\cdot p_{+0}}( 4M^2_\pi-7p_{+0}^2+6 k \cdot p_{+0}) +2\frac{k\cdot p_+}{k\cdot p_{+0}} \frac{p_{+0}^2}{k\cdot p_{+0}}(p_{+0}^2-4M^2_\pi) \bigg] \nonumber\\ && +{\bar R}(p_{+0}^2,k \cdot p_{+0})\big[-M^2_\pi+2 k\cdot p_+ (\frac{p_{+0}^2}{k\cdot p_{+0}}-1)-\frac{1}{3}(\alpha-1)M^2_\pi\big] \nonumber\\ && +\frac{1}{24\pi^2} k\cdot p_+ (1-\frac{p_{+0}^2}{k\cdot p_{+0}}) + \frac{1}{96 \pi^2} p_{+0}^2\bigg\}\ , \end{eqnarray} \vs{2}\begin{equation} {\cal A}^C_5(k,k';p_0,p_+,p_-)= -\frac{e^2}{24\pi^2F_\pi^5}\bigg\{ {\bar J}\big((k'-p_{-})^2\big)\big[4M^2_\pi-(k'-p_{-})^2\big] -{\bar J}(p_{+0}^2)(4M^2_\pi-p_{+0}^2) -\frac{1}{24\pi^2}k\cdot p_{+0}\bigg\}\ , \\[2mm]\end{equation} \begin{eqnarray} {\cal A}^C_6(k,k';p_0,p_+,p_-)&=& -\frac{e^2}{4\pi^2F_\pi^5}\bigg\{ \frac{1}{6} \big[\bar{J}\big((k'-p_{-})^2\big)-\bar{J}(p_{+0}^2)\big] \bigg[p_{+0}^2-4M^2_\pi -6 k\cdot p_+ \nonumber\\ &&\qquad+4 \frac{k\cdot p_+}{k\cdot p_{+0}} (p_{+0}^2- M^2_\pi) +\frac{k\cdot p_+}{(k\cdot p_{+0})^2}p_{+0}^2 (4M^2_\pi-p_{+0}^2)\bigg] \nonumber\\ && -\frac{1}{3}\bar{J}\big((k'-p_{-})^2\big)\big[ k\cdot p_0+2 M^2_\pi\frac{k\cdot p_+}{k\cdot p_{+0}}\big] \nonumber\\ && +{\bar R}\big(p_{+0}^2,(k-p_{+0})^2\big) (k \cdot p_+) \big[2-\frac{p_{+0}^2} {k\cdot p_{+0}}\big] \nonumber\\ && +\frac{1}{32 \pi^2} \frac{k\cdot p_+}{k\cdot p_{+0}}(k'-p_-)^2 +\frac{1}{96\pi^2}\bigg(\frac{4}{3} k\cdot p_+- p_{+0}^2 \frac{k\cdot p_+} {k\cdot p_{+0}}\bigg) \nonumber\\ && +\frac{1}{144 \pi^2} k\cdot p_{+0}\bigg\}\ . \end{eqnarray} The amplitude $A_1^C$ again contains a contribution with a pole at $s=M_\pi^2$, which is given by the product of the ${\cal O}({\mbox p}^6)$ $\pi^0\to\gamma\gamma$ amplitude ${\cal A}^{\pi^0\to\gamma\gamma}$ times the (off--shell) $\pi^0\pi^0\to\pi^+\pi^-$ amplitude ${\cal A}^{00;+-}$, with \begin{eqnarray} -{\cal A}^{00;+-}(p_{+-}^2,p_{+0}^2,p_{-0}^2) &&= \ \frac{\beta}{F_\pi^2}(p_{+-}^2 -\frac{4}{3} M_\pi^2)+\frac{1}{3F_\pi^2}\alpha M_\pi^2\nonumber\\ &&+{{\lambda_1}\over{F_\pi^4}}(p_{+-}^2-2M_\pi^2)^2 +{{\lambda_2}\over{F_\pi^4}}[(p_{+0}^2-2M_\pi^2)^2+(p_{0-}^2-2M_\pi^2)^2] \nonumber\\ &&+{1\over{6F_\pi^4}} \bar{J}(p_{+-}^2)\bigg[4(p_{+-}^2-\frac{4}{3}M_\pi^2 +\frac{5}{6}\alpha M_\pi^2)^2-(p_{+-}^2-\frac{4}{3}M_\pi^2-\frac{2}{3} \alpha M_\pi^2)^2 \bigg] \nonumber\\ &&+{1\over{12F_\pi^4}} \bar{J}(p_{+0}^2)\bigg[3(p_{+0}^2-\frac{4}{3}M_\pi^2 -\frac{2}{3}\alpha M_\pi^2)^2+(p_{+-}^2-p_{-0}^2)(p_{+0}^2-4M_\pi^2) \bigg] \nonumber\\ &&+{1\over{12F_\pi^4}} \bar{J}(p_{-0}^2)\bigg[3(p_{-0}^2-\frac{4}{3}M_\pi^2 -\frac{2}{3}\alpha M_\pi^2)^2+(p_{+-}^2-p_{+0}^2)(p_{-0}^2-4M_\pi^2) \bigg]\ , \nonumber\\ \end{eqnarray} where $p_{+-}^2+p_{+0}^2+p_{-0}^2=s+3M_\pi^2$. The remaining parameters $\gamma_{+-}$ and $\gamma_{+-}'$ which appear in the amplitudes ${\cal A}_1^C$, ... ${\cal A}_6^C$ contain the contributions from the low--energy constants and chiral logarithms~: \begin{eqnarray} \gamma_{+-}&=& -{\widehat m}^2(a_2+2b_2+4c_1) \nonumber\\ &&+{{M_\pi^2}\over{96\pi^2F_\pi^2}}\ln{{M_\pi^2}\over{\mu^2}} (\alpha-1) +{{M_\pi^2}\over{96\pi^2F_\pi^2}} (\alpha-\frac{7}{3})+{\widehat m}^2\delta\gamma_{+-}\ ,\lbl{g} \end{eqnarray} \begin{eqnarray} \gamma_{+-}' &=& 8{\widehat m}(2A_{12}-A_{13}) + 8M_\pi^2(A_7-A_8)+6{\widehat m}^2a_3\nonumber\\ && -{{M_\pi^2}\over{32\pi^2F_\pi^2}}\ln{{M_\pi^2}\over{\mu^2}} +{\widehat m}^2\delta\gamma_{+-}'\ .\lbl{g'} \end{eqnarray} With the expressions given above, it is straightforward to check that the three conditions listed before Eq. \rf{iso1} as well as Eq. \rf{iso2} are satisfied, which provides a non--trivial check of our calculation. The isospin relation \rf{iso1} implies that the condition \vs{2}\begin{equation} 2 + 3\gamma_{+-}+\gamma_{00}+{{M_\pi^2}\over{24\pi^2F_\pi^2}} ={\widehat m}^2( 2a_3 - 3t'') \\[2mm]\end{equation} must hold. This requires that the contributions of the ${\cal L}_{(4,2)}^-$ counterterms to $\gamma_{+-}$, which we have denoted as ${\widehat m}^2\delta\gamma_{+-}$ in the expression \rf{g}, have to satisfy \vs{2}\begin{equation} 3t''+3\delta\gamma_{+-}=0\ .\lbl{isocond} \\[2mm]\end{equation} The contributions of the ${\cal L}_{(4,2)}^-$ counterterms to $\gamma_{+-}'$, ${\widehat m}^2\delta\gamma_{+-}'$ in the expression \rf{g'}, are not constrained by isospin symmetry. Upon using the information provided by Table~1, it is straightforward to check that $\alpha$, $\beta$ and $\gamma_{00}$ are scale independent by themselves (the parameters $\lambda_1$ and $\lambda_2$ were directly expressed in terms of the scale independent quantities ${\bar l}_1$ and ${\bar l}_2$ defined in Ref.~\cite{GL1}). In order that the amplitudes ${\cal A}^{\pi^0\to\gamma\gamma}$, ${\cal A}^N$ and ${\cal A}^C$ be independent of the subtraction scale $\mu$, the parameters $t$, $t'$, $t''$, $\gamma_{+-}$ and $\gamma_{+-}'$ must be separately scale independent. This requires that $\delta\gamma_{+-}$ and $\delta\gamma_{+-}'$ themselves are scale independent. Since we have not worked out the structure of the one--loop divergences of the generalized generating functional in the odd intrinsic parity sector, we could not perform these checks explicitly. >From the above formulae, one may infer the expressions of the amplitudes in the standard case \cite{pere96,perethesis}. Since only the result for the neutral amplitude ${\cal A}^N$ was displayed explicitly in Ref.~\cite{pere96}, and the expressions of the amplitudes ${\cal A}^C_i$ in the charged channel are only available from the unpublished work \cite{perethesis}, we describe in some detail the necessary steps to obtain them. Their general structure is of course unchanged, the differences occur only in the expressions of the various combinations of low--energy constants that are involved. In particular, the contributions from ${\cal L}_{(0,3)}$, ${\cal L}_{(2,2)}$, ${\cal L}_{(0,4)}$, and ${\cal L}^-_{(4,2)}$ are relegated to higher orders. For the remaining constants, the correspondance with the usual SChPT notation is given as follows \begin{eqnarray} \hat m \xi^{(2)}_{\mbox{\scriptsize st}} &=& \frac{1}{16\pi^2 F_\pi^2}(\bar l_4+ \ln \frac{M_\pi^2} {\mu^2})\ , \nonumber \\ \alpha_{\mbox{\scriptsize st}} &=& 1 + \frac{M_\pi^2}{32 \pi^2 F_\pi^2}(4 \bar l_4 -3 \bar l_3 -1)\ , \nonumber \\ \beta_{\mbox{\scriptsize st}} &=& 1 + \frac{M_\pi^2}{8 \pi^2 F_\pi^2}( \bar l_4 -1)\ , \nonumber \\ \gamma_{00,{\mbox{\scriptsize st}}}&=& 0\ ,\nonumber \\ \gamma_{+-,{\mbox{\scriptsize st}}}&=& - \frac{M_\pi^2}{72 \pi^2 F_\pi^2}\ , \nonumber \\ \gamma_{+-,{\mbox{\scriptsize st}}}'&=& 8M_\pi^2 (A_7-A_8+ \frac{2 A_{12}-A_{13}}{2B})- \frac{M_\pi^2}{32 \pi^2 F_\pi^2} \ln \frac{M_\pi^2}{\mu^2}\ . \lbl{Sval} \end{eqnarray} We have checked that upon substituting these expressions into the one--loop amplitudes ${\cal A}^N$ and ${\cal A}^C$ given above, we recover the results of the standard case, up to the contributions from the counterterms contained in $t$ and $t'$, which were not included in Refs.~\cite{pere96,perethesis}. \section{Counterterm Estimates and Numerical Results} \label{COUNTER} \setcounter{equation}{0} In order to make numerical estimates for the cross sections based on the ${\cal O}({\mbox p}^6)$ amplitudes, we first need to fix or estimate the values of the various counterterms involved. \noindent ~~{\bf i)}~$\underline{ \lambda_1,\ \lambda_2}$~: At order ${\cal O}({\mbox p}^4)$, these parameters are related to ${\bar l}_1$ and ${\bar l}_2$ through Eq.~\rf{l1l2}. The values of these low--energy constants in the standard case have been the subject of numerous studies in the literature \cite{GL1,l1l2Kl4,l1l2coll,BCEGS,girlanda97}. A first determination at order ${\cal O}({\mbox p}^4)$ was given in Ref.~\cite{GL1}, using information from the D--wave $\pi$--$\pi$ scattering lengths. The corresponding values, taken from a recent numerical re--analysis~\cite{l1l2G}, are \vs{2}\begin{equation} {\bar l}_{1,{\mbox{\scriptsize GL}}} = -2.15\pm 4.30\ , \ \ {\bar l}_{2,{\mbox{\scriptsize GL}}} = 5.84\pm 1.72\ , \\[2mm]\end{equation} leading to \vs{2}\begin{equation} \lambda_{1,{\mbox{\scriptsize GL}}} = (-7.35\pm 9.06)\times 10^{-3}\ , \ \ \lambda_{2,{\mbox{\scriptsize GL}}} = (10.57\pm 3.63)\times 10^{-3}\ .\lbl{l1l2G} \\[2mm]\end{equation} The parameters $\lambda_1$ and $\lambda_2$ can also be determined directly, {\it via} a set of rapidly convergent sum--rules \cite{pipipaper2}, from the knowledge, at order two loops, of the $\pi$--$\pi$ scattering amplitude $A(s\vert t,u)$ in GChPT \cite{pipipaper1}, and from medium energy data on $\pi$--$\pi$ phase shifts. The values obtained this way correspond to a determination at order ${\cal O}({\mbox p}^6)$. They depend only very weakly on the values of the parameters $\alpha$ and $\beta$ when the latter are varied within the ranges specified below, and read \vs{2}\begin{equation} \lambda_{1} = (-6.1\pm 2.2)\times 10^{-3}\ , \ \ \lambda_{2} = (9.6\pm 0.5)\times 10^{-3}\ .\lbl{l1l2K} \\[2mm]\end{equation} These values are compatible with those given in eq.~\rf{l1l2G}, but are affected by much smaller error bars. The analysis may even be refined in the standard case, using the information on the SChPT two--loop $\pi$--$\pi$ amplitude obtained in Ref.~\cite{BCEGS}, leading to the following values~\cite{girlanda97}, \vs{2}\begin{equation} \lambda_{1,{\mbox{\scriptsize st}}} = (-5.7\pm 2.2)\times 10^{-3}\ , \ \ \lambda_{2,{\mbox{\scriptsize st}}} = (9.3\pm 0.5)\times 10^{-3}\ . \lbl{l1l2Kst} \\[2mm]\end{equation} \noindent ~~{\bf ii)}~$\underline{\alpha,\ \beta}$~: At leading order, $\alpha$ is directly correlated to the size of the condensate, see Eq.~\rf{alphalead}. As such, the value of $\alpha$ is not predicted by GChPT, but remains a free parameter, that can {\it a priori} be varied in the range $1\stackrel{\displaystyle <}{{}_{\displaystyle\sim}}\alpha\stackrel{\displaystyle <}{{}_{\displaystyle\sim}} 4$. At order ${\cal O}({\mbox p}^4)$, the relationship between $\alpha$ and the ratio $2{\widehat m} B/M_{\pi}^2$ becomes more complicated, as shown in Eq.~\rf{alphaexp}. The corresponding next--to--leading and next--to--next--to--leading corrections have been estimated in Ref.~\cite{pipipaper1} (see in particular the figures 8 and 10 in that reference). Notice that the analysis of Ref.~\cite{pipipaper1} was done within the framework of $SU(3)_{\mbox{\scriptsize L}}\times SU(3)_{\mbox{\scriptsize R}}$ chiral perturbation theory. Working with only two light flavours as in the present paper might further reduce the uncertainties in the correspondance between the value of $\alpha$ and the size of the condensate, due to the lower number of unknown counterterms involved, and due to the absence of large contributions from the chiral logarithms induced by the kaon loops \cite{girl98}. However, the results of Ref.~\cite{pipipaper1} are sufficient for our present purposes, and we shall not pursue that matter further. Once $\alpha$ is given, the parameter $\beta$ is also constrained by low--energy $\pi$--$\pi$ data. The correlation between $\alpha$ and $\beta$, which results from the Morgan--Shaw universal curve~\cite{morganshaw}, has also been studied beyond leading order in \cite{pipipaper1}, and is summarized in Fig.~6 of that reference. For the subsequent numerical analyses, we shall take the values given in Table 2 below. The values given in the second line of this table correspond to the one--loop values $\alpha_{\mbox{\scriptsize st}}$ and $\beta_{\mbox{\scriptsize st}}$ of the standard case, which follow from the expressions given in Eq.~\rf{Sval}, and from the values ${\bar l}_3 = 2.9\pm 2.4$ \cite{GL1}, ${\bar l}_4 = 4.4\pm 0.3$ \cite{pere98}. Two points are worth being remembered. The first is that, independently of the value of $\alpha$, $\beta$ stays close to unity. The second point is that the standard case allows to make a very precise prediction for the value of $\alpha$ at order ${\cal O}({\mbox p}^4)$, {\it viz.} $\alpha_{\mbox{\scriptsize st}} = 1.06\pm 0.06$. Furthermore, this value is barely affected by the corrections at order ${\cal O}({\mbox p}^6)$~: The analysis of Ref.~\cite{girlanda97}, based on the results of \cite{BCEGS}, gives $\alpha_{\mbox{\scriptsize st}}=1.07\pm 0.01$ and $\beta_{\mbox{\scriptsize st}}=1.105\pm 0.015$ at next--to--next--to--leading order. Therefore, any significant deviation of the value of $\alpha$ from unity would provide evidence for a departure from the standard scenario of chiral symmetry breaking with a strong condensate. \noindent ~~{\bf iii)}~$\underline{ t,\ t',\ t''}$~: The constants $t$, $t'$ and $t''$ appear in the expression of the $\pi^0\to\gamma\gamma$ amplitude \rf{ggpi0}. The uncertainty on the experimental value of the decay rate~\cite{PDG}, $\Gamma({\pi^0\to\gamma\gamma})=7.74\pm 0.56$ eV, only yields a very weak constraint on the combination that appears in ${\cal A}^{\pi^0\to\gamma\gamma}$, {\it viz.} ${\widehat m} t +M_\pi^2 t' +{\widehat m}^2 t'' = (0.0\pm 3.6)\times 10^{-2}$. This is comparable to the estimate one would obtain through naive dimensional analysis \cite{dimann}. In addition, as shown in Ref.~\cite{bachir95}, isospin breaking effects can be sizeable in ${\cal A}^{\pi^0\to\gamma\gamma}$. Further information may be obtained by making use of the sum--rules considered in Ref.~\cite{bachir95}. The corrections due to $t'$ were however not taken into account there, but the analysis of \cite{bachir95} is easily extended to the more general situation. In the generalized case, a similar set of sum--rules can be established, but they do not yield complete information on the three constants $t$, $t'$ and $t''$. We have summarized this analysis in the Appendix for the interested reader. Here, we only quote the values that we shall use in the sequel (the estimate for $t''$ follows from naive dimensional analysis), \vs{2}\begin{equation} {\widehat m} t\,=\, (6\pm 12)\times 10^{-3}\, , \ \ M_\pi^2t'\,=\,(-3\pm 3)\times 10^{-3}\, , \ \ {\widehat m}^2 t''\,\sim\,\pm 1\times 10^{-3} \ .\lbl{genWZresult} \\[2mm]\end{equation} \noindent ~~{\bf iv)}~$\underline{ \gamma_{00},\ \gamma_{+-},\ \gamma_{+-}'}$~: The main difficulty in obtaining numerical estimates for these parameters comes from the lack of knowledge, in the generalized case, on the contributions from the low--energy constants of ${\cal L}^+_{(2,2)}$. In the spirit of Vector Meson Dominance (VMD), we have estimated the contributions of vector mesons to the two amplitudes ${\cal A}^N$ and ${\cal A}^C$, as done in Ref.~\cite{pere96} for the standard case. This way, we find that $t$, $t'$ and $t''$ receive no contribution, whereas the contribution to $\gamma_{00}$, $\gamma_{+-}$ and $\gamma_{+-}'$ read \vs{2}\begin{equation} \gamma_{00}\vert_{\mbox{\scriptsize VMD}}\,=\, \gamma_{+-}\vert_{\mbox{\scriptsize VMD}}\,=\,0\ , \ \ \gamma_{+-}'\vert_{\mbox{\scriptsize VMD}}\,=\, -\frac{3}{4}\frac{M_\pi^2}{M_V^2}\,\sim\, -0.024\ .\lbl{VMDresult} \\[2mm]\end{equation} We take \rf{VMDresult} as the values of these constants at the scale $\mu\sim M_V$=770 MeV, and estimate the error associated to the VMD approximation and to the presence of the low-energy constants from ${\cal L}_{(2,2)}^+$ in the expressions of $\gamma_{00}$ and of $\gamma_{+-}$ by varying the scale $\mu$ of the corresponding chiral logarithms between 500 MeV and 1 GeV. The resulting values for $\gamma_{00}$ and $\gamma_{+-}$ are shown in Table 2. In the case of $\gamma_{+-}'$, we obtain a constant value which, to a very good precision, is compatible with zero. \begin{table*}[htbp] \begin{center} \begin{tabular}{|c|c|c|c|} \hline $\alpha$ & $\beta$ & $\gamma_{00}\times 10^3$ & $\gamma_{+-}\times 10^3$ \\ \hline $1.06\pm 0.06$ & $ 1.103 \pm 0.008$ & $0$ & $-3$ \\ \hline 1.5 & $1.06 \pm 0.06$ & $8\pm 3$ & $-5.7\pm 0.8$ \\ \hline 2 & $1.07 \pm 0.06$ & $16\pm 5$ & $-8.4\pm 1.6$ \\ \hline 2.5 & $1.08 \pm 0.06$ & $24\pm 8$ & $-11.0\pm 2.4$ \\ \hline 3 & $1.11 \pm 0.06$ & $32\pm 10$ & $-13.7\pm 3.2$ \\ \hline \end{tabular} \end{center} \begin{center} {\bf Table 2:}{\it Values of $\beta$, $\gamma_{00}$ and $\gamma_{+-}$ for different values of $\alpha$.} \end{center} \end{table*} \indent With the above inputs at hand, we may now consider a few numerical applications. In Fig.~2, we have plotted the tree--level and one--loop cross sections for the charged channel, $\sigma^C_{\mbox{\scriptsize tree}}(s,\alpha)$ and $\sigma^C(s,\alpha)$, in the threshold region, $3 M_\pi \le \sqrt{s} \le 0.5$~GeV, where we expect the one--loop expression of the amplitude to be reliable, and for different values of $\alpha$. We have used $M_{\pi^\pm} = M_{\pi^0}= 135$~MeV in the amplitude and the experimental values, as quoted in \cite{PDG}, in the phase space integrals. Let us first concentrate on the standard case ($\alpha\sim 1$) which has been discussed before in the literature. In the second half of the energy region that we have shown, the correction as compared to the tree level cross section amounts approximatively to a factor of two, if we consider the central values. This agrees with the unpublished result \cite{perethesis}, but is much less than previously found in \cite{pere96}. On the other hand, the error bars induced by the uncertainties attached to the various conterterm contributions that enter the ${\cal O}({\mbox p}^6)$ amplitude are important. A closer analysis (see also below) reveals that the main contribution comes from the uncertainty on the value of $\lambda_1$ given in \rf{l1l2Kst}. This shows also that the sensitivity of the cross section on the ${\cal O}({\mbox p}^6)$ counterterms becomes already sizeable even at such low energies. Coming now to the dependence with respect to $\alpha$, we see that here also the higher order corrections have a deep influence and upset the situation that prevailed at tree level. For $\alpha\ge 2$ the one--loop cross sections are suppressed as compared to their tree--level values, and become even smaller than $\sigma^C(s,1)$ as the energy increases. Unfortunately, the present theoretical error bars make it difficult to disentangle the different situations in practice from the knowledge of the total cross section alone. \indent \centerline{\psfig{figure=fig2.gg3pi.ps,height=12.5cm,angle=-90}} \noindent {\bf Figure 2} : {\it The cross section $\sigma(s,\alpha)$ (in logarithmic scale) for $\gamma\gamma\to\pi^+\pi^-\pi^0$ as a function of the center of mass total energy, for three values of $\alpha$. Also shown are the corresponding curves for the tree level cross section {\rm $\sigma_{\mbox{\scriptsize tree}}(s,\alpha)$}.} \indent The cross sections $\sigma^N_{\mbox{\scriptsize tree}}(s,\alpha)$ and $\sigma^N(s,\alpha)$ in the neutral channel have been plotted in Fig.~3. Whereas $\sigma^N_{\mbox{\scriptsize tree}}(s,\alpha)\sim \alpha^2$, the corrections are seen to have an even more drastic influence on the behaviour of the cross section as a function of energy than in the charged channel. Unfortunately, as far as the dependence on $\alpha$ is concerned, the picture is again totally blurred by the uncertainties, which, for the sake of clarity, we have not shown, but which are even more important than in the charged case. In both channels, the origin of the large error bars is a consequence of the highly destructive interferences between the various amplitudes. In the charged case, this interference was already present at tree level, and is even accentuated by the loop effects. In the neutral case, the strong $\alpha$ dependence of the single amplitude that contributes at lowest order is, in a similar way, washed out by the interferences between the two one--loop amplitudes ${\cal A}_1^N$ and ${\cal A}_2^N$. In order to illustrate this point, we show, in Fig.~4, how the neutral cross section looks like if only the amplitude ${\cal A}_1^N$ is considered. The importance of the interference effects with the second amplitude ${\cal A}_2^N$ appears clearly upon comparing Figs.~3 and 4 (for the standard case, a similar observation was already made by the authors of Ref.~\cite{pere96}). Unfortunately, we have not found a simple way to extract the contribution from ${\cal A}_1^N$ only~: for photons with perpendicular polarizations, the two amplitudes contribute, and the destructive interference between them is again at work, while for photons with parallel polarizations, only the very small contribution from ${\cal A}_2^N$, which vanishes at tree level, is singled out. We have tried to investigate whether looking at more refined observables allows to reach better perspectives from this point of view. \indent \centerline{\psfig{figure=fig3.gg3pi.ps,height=12.5cm,angle=-90}} \noindent {\bf Figure 3} : {\it The cross section $\sigma(s,\alpha)$ (in logarithmic scale) for $\gamma\gamma\to\pi^0\pi^0\pi^0$ as a function of the center of mass total energy, for three values of $\alpha$. Also shown are the corresponding curves for the tree--level cross section {\rm $\sigma_{\mbox{\scriptsize tree}}(s,\alpha)$}.} \indent \noindent We have, for instance, looked at the invariant mass distribution of the two charged pions in the $\gamma\gamma\to\pi^+\pi^-\pi^0$ channel. The result is shown in Fig.~5. As one may observe, the error bars are much less important than for the total cross section, and the different values of $\alpha$ can be distinguished over a substantial portion of the energy range that has been considered. Actually, analyses of this type usually require sufficiently high statistics, which also represents a problem in the present case. \indent \centerline{\psfig{figure=fig4.gg3pi.ps,height=12.5cm,angle=-90}} \noindent {\bf Figure 4} : {\it The cross section $\sigma_{A_1^N}(s,\alpha)$ for $\gamma\gamma\to\pi^0\pi^0\pi^0$ obtained by taking into account the contribution from the amplitude $A_1^N$ alone, for different values of $\alpha$.} \indent Indeed, in both channels, the cross sections are very small at low energies, orders of magnitude below, for instance, the corresponding cross sections for the $\gamma\gamma\to\pi\pi$ processes. We have therefore also estimated the numbers of events that could be expected at an $e^+-e^-$ collider for the two--photon total invariant mass $\sqrt{s}$ below a maximal energy $E_{\mbox{\scriptsize max}}$. As typical examples, we have considered two instances of symmetric $e^+-e^-$ colliders. The first case corresponds to the Daphne $\phi$--factory~\cite{daphne}, with a total beam energy of $E_{\mbox{\scriptsize beam}}\,=\,510$ MeV, and a nominal integrated luminosity of $5\times 10^6$ nb$^{-1}$ per year. The second case concerns a $\tau$--Charm Factory configuration~\cite{tauCF}, with a beam energy four times as large as for Daphne, and a design integrated luminosity of $10^7$ nb$^{-1}$ per year. The numbers of events are obtained upon convoluting the above cross sections with the corresponding photon luminosities quoted in \cite{Lia} (since we are not interested in resonant $\eta$ production, whenever necessary we avoided the $\eta$ peak by applying a cut on $m_{\gamma \gamma}$ such that only the events with $m_{\gamma \gamma}< M_\eta -\Delta$ or $m_{\gamma \gamma} > M_\eta +\Delta$, with $\Delta=20~$MeV are accepted), \vs{2}\begin{equation} {d L_{\gamma\gamma}\over d m_{\gamma\gamma}}= {4\over m_{\gamma\gamma}}\big({\alpha\over \pi} \ln {E_{\mbox{\scriptsize beam}}\over m_e}\big)^2 \big[-(2+z^2)^2 \ln z -(1-z^2) (3+z^2) \big], \\[2mm]\end{equation} where $m_{\gamma\gamma}=\sqrt{s}$ is the photon--photon center of mass energy, $z=m_{\gamma\gamma}/2E_{\mbox{\scriptsize beam}}$, and $m_e$ is the electron mass. For the luminosity quoted above, the expected total number of $\gamma\gamma\to\pi^+\pi^-\pi^0$ events per year at energies $\sqrt{s}\stackrel{\displaystyle <}{{}_{\displaystyle\sim}} 0.6$~GeV is around $5\pm 1$ for Daphne (independently of $\alpha$) and even less in the neutral case. In the case of a $\tau$--Charm Factory, the total number of events becomes sizeable, and the results are shown, for the two modes, in the fourth column of Table~3, \indent \centerline{\psfig{figure=fig5.gg3pi.ps,height=12.5cm,angle=-90}} \noindent {\bf Figure 5} : {\it The histogram of the distribution of the invariant mass $m_{\pi^+\pi^-}$ of the two final charged pions in the process $\gamma\gamma\to\pi^+\pi^-\pi^0$ for different values of $\alpha$ and $\sqrt{s} \le E^{max} = 0.6$~GeV. Also shown are the curves correponding to the same distribution, but from the lowest order amplitude alone.} \indent \noindent for different choices of $E_{\mbox{\scriptsize max}}$ (second column) and for different values of $\alpha$ (third column). Also shown are the corresponding uncertainties $\Delta$(\#events) (fifth column), while the remaining entries show various sources of contributions to the total error $\Delta$(\#events). The latter was obtained upon adding the individual contributions in quadrature. As mentioned above, the main source of error comes from the uncertainty on the value of $\lambda_1$. A sizeable but not necessarily drastic reduction of the latter would already allow to distinguish the standard case of a strong condensate ($\alpha\sim 1$) from situations where spontaneous breakdown of chiral symmetry would be triggered by a much weaker condensate ($\alpha \stackrel{\displaystyle >}{{}_{\displaystyle\sim}} 2$). \indent \noindent \begin{table*}[t] \begin{center} \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline Mode & $E_{\mbox{\scriptsize max}}({\mbox GeV})$ & $\alpha$ & \#events & $\Delta$(\#events) & $\Delta\lambda_1$ & $\Delta\lambda_2$ & $\Delta\beta$ & $\Delta{\widehat m} t$ \\ \hline & & 1 & 16 & 4 & 3.3 & 1.3 & 0.3 & 1.3 \\ &0.50 & 2 & 11 & 1 & 1 & 0.2 & 0.2 & 0.4 \\ & & 3 & 12 & 2 & 1 & 0.6 & 1.8 & 0.5 \\ \cline{2-9} & & 1 & 56 &11 & 9.7 & 3.6 & 0.6 & 3.5 \\ $\pi^+\pi^-\pi^0$ &0.55 & 2 & 40 & 5 & 4.5 & 1.3 & 0.1 & 1.8 \\ & & 3 & 38 & 3 & 0.2 & 0.6 & 3.1 & 0.1 \\ \cline{2-9} & & 1 &327 &56 &50 & 18 & 1.8 & 14 \\ &0.60 & 2 &273 &39 &35 & 11 & 2.7 & 10 \\ & & 3 &246 &26 &24 & 5.7 & 4.4 & 7.3 \\ \hline \hline & & 1 & 14 & 7 & 5.9 & 2.7 & 0.7 & 2.3 \\ &0.50 & 2 & 1 & 2 & 1.2 & 0.6 & 1 & 0.5 \\ & & 3 & 5 & 4 & 2.6 & 1.2 & 2.6 & 1.3 \\ \cline{2-9} & & 1 & 42 &18 &15.8 & 7.2 & 1.8 & 5.6 \\ $\pi^0\pi^0\pi^0$ &0.55 & 2 & 8 & 9 & 5.9 & 2.7 & 4.8 & 2.1 \\ & & 3 & 6 & 3 & 1.7 & 0.8 & 2.0 & 1.0 \\ \cline{2-9} & & 1 &211 &81 &71 & 32 & 6.3 & 19 \\ &0.60 & 2 &91 &59 &45 & 20 & 28 & 12 \\ & & 3 &51 &34 &26 & 12 & 15 & 6.6 \\ \hline \end{tabular} \end{center} {\bf Table 3:}~{\it Number of events for $\gamma\gamma\to\pi^+\pi^-\pi^0$ and for $\gamma\gamma\to\pi^0\pi^0\pi^0$ at a $\tau$--Charm Factory as a function of the maximal energy $E_{\mbox{\scriptsize max}}({\mbox GeV})$, and the principal sources of error.} \end{table*} \section{Summary and Conclusions} \label{conclu} In the present paper, the amplitudes of the processes $\gamma\gamma\to \pi^0\pi^0\pi^0$ and $\gamma\gamma\to \pi^+\pi^-\pi^0$ have been computed in the framework of $SU(2)_{\mbox{\scriptsize L}}\times SU(2)_{\mbox{\scriptsize R}}$ Generalized Chiral Perturbation Theory to ${\cal O}({\mbox p}^6)$ precision. The corresponding generating functional has been constructed explicitly in Section 3, and the structure of its divergences in the sector of even intrinsic parity has been analysed. The resulting amplitudes, worked out in Section~4, satisfy the isospin relations that we have established in Section 2 (to the best of our knowledge, these relations have not been discussed previously in the literature). When restricted to the standard case, specified by the choice of parameters as indicated in Eq.~\rf{Sval}, we recover the results obtained by previous authors \cite{pere96,perethesis}, both in the neutral and in the charged case. Finally, we have estimated the counterterms that enter the one-loop amplitudes and we have performed some numerical analyses in Section 5. We have, in particular, shown that the error bars associated to the counterterm estimates become important, especially in the neutral channel, as a consequence of highly destructive interference effects between the various amplitudes that build up the total cross sections. We have also considered the possible detection of these processes at Daphne and at a $\tau$--Charm Factory. Unfortunately, and precisely because of these large interference effects, the expected number of events is rather discouraging in the first case. Depending on the actual values of the counterterms and on $\alpha$, it is hard to expect more than $\approx 5$ events per year with total invariant mass lower than $500$~MeV. The number of events increases substantially when allowing larger invariant masses, but at the expense of working in an energy region where the ${\cal O}(p^6)$ expressions are probably less reliable, since higher order terms can become important. The computation at ${\cal O}({\mbox p}^8)$ would allow a better control of the cross sections at larger momenta, thus allowing a substantial increase of the number of events already at Daphne. The required amount of work seems excessive, though, and would make sense only if conducted in parallel with a better determination of ${\bar l}_1$ (for instance, from a two--loop analysis of $K_{\ell 4}$ decays), by far the main source of theoretical uncertainties at present. We rather expect interesting and realistic prospects in this field to come from future machines, like the $\tau$--Charm Factory, which run at higher energy and higher luminosity. \section*{Acknowledgments} We are pleased to thank A. Bramon for discussions and for his interest and collaboration at the early stages of this work, and J.~Bijnens for discussions. One of us (M.K.) wishes to acknowledge clarifying correspondance and/or discussions with I.~Kogan and J.~Stern on the content of Ref.~\cite{KKS98}, and L.~Girlanda for discussions and for sharing information on unpublished work. He also thanks the Universitat Polit\`ecnica de Catalunya and the Universitat Aut\`onoma de Barcelona for their hospitality. Ll.~A. thanks the Institut de Physique Nucl\'eaire in Orsay for its hospitality. P.~T. thanks FEN department, where most of his contribution to the present work was done. This work has been partially supported by the TMR Program, EC--Contract N. CT98--0169. Ll.~A. and P.~T. received partial support from the CICYT research project AEN95--0815, while the work of J.~K. and of P.~T. was supported by the Schweizerischer Nationalfonds and by the Swedish Research Council (NFR), respectively. \indent \indent \noindent {\Large{\bf Appendix}} \renewcommand{\theequation}{A.\arabic{equation}} \setcounter{equation}{0} In this Appendix, we give a brief description of our analysis of the anomalous counterterms $A_2$, $A_3$, $A_4$ and $A_6$ which is based on the approach of Ref.~\cite{bachir95}. The starting point are the invariant amplitudes $\Pi_{VVP}(p^2,q^2,r^2)$ and $\Pi_{VVP}^0(p^2,q^2,r^2)$ of the vector--vector--pseudoscalar three--point correlation functions in the three flavour chiral limit (we take the definition given by Eqs.~(3) and (4) of \cite{bachir95}). At $p^2=q^2=0$, the loop contribution vanish, and one has (notice the absence of the pion pole in the second equality) \begin{eqnarray} \Pi_{VVP}(0,0,r^2) &=& \frac{2B_0N_c}{16\pi^2r^2}\,+\,\frac{1}{4\pi^2} \big[ 8A_4-16B_0(A_2-2A_3)\big]\ ,\nonumber\\ \Pi_{VVP}^0(0,0,r^2) &=& \frac{1}{4\pi^2}\big[ 8A_4 +24 A_6\big]\ . \end{eqnarray} In the chiral limit, the counterterms from ${\cal L}_{(4,2)}^-$ do not contribute, so that the above result holds both in SChPT and in GChPT. Although the contributions from the low--energy constants $A_2$ and $A_3$ from ${\cal L}_{6,0}^-$ were omitted in \cite{bachir95}, one may follow the same steps as described there (we have also kept the same notation), and end up with the following set of sum--rules~: \vs{2}\begin{equation} \frac{1}{4\pi^2} \big[ 8A_4-16B_0(A_2-2A_3)\big] \,=\, -\frac{B_0}{2M_V^2}\bigg\{ \frac{F_0^2}{M_V^2}\,+\,\frac{N_c}{4\pi^2} \left(\frac{M_V}{M_P}\right)^2{\mbox{tan}}\Theta\frac{{\cal A}(\pi'\to\gamma\gamma)}{{\cal A}(\pi\to\gamma\gamma)}\bigg\}\ ,\lbl{bachir1} \\[2mm]\end{equation} and \vs{2}\begin{equation} \frac{1}{4\pi^2} \big[ 24A_6+16B_0(A_2-2A_3)\big] \,=\,-\frac{3}{8} \frac{B_0G_{\eta'}}{M_{\eta '}^2}\sqrt{6}{\cal A}(\eta'\to\gamma\gamma)\ . \lbl{bachir2} \\[2mm]\end{equation} In addition, one needs to know the expression of the amplitude of the two--photon decay of the $\eta$, ${\cal A}(\eta\to\gamma\gamma )=-t_1(k,\epsilon,k',\epsilon'){\cal A}^{\eta\to\gamma\gamma}$, which, in analogy to the $\pi^0\to\gamma\gamma$ amplitude ${\cal A}^{\pi^0\to\gamma\gamma}$ of Eq.~\rf{ggpi0}, may be written as \vs{2}\begin{equation} {\cal A}^{\eta\to\gamma\gamma}\,=\,\frac{e^2}{4\sqrt{3} \pi^2 F_\pi} \big[\frac{F_\pi}{F_\eta} +\frac{1}{3}{\widehat m} (5-2r)t+ \frac{128}{3}{\widehat m} (1-r)A_6+M_\eta^2t'+{\widehat m}^2{\widetilde t}''(r)+\frac{2}{3}{\widehat m}^2(1+2r^2)a_3\big]\ .\lbl{ggeta} \\[2mm]\end{equation} In this last expression, $t$ and $t'$ are related to $A_2$, $A_3$ and $A_4$ according to Eq.~\rf{ttprime}, while $r$ stands for the quark mass ratio $m_s/{\widehat m}$, and ${\widehat m}^2{\widetilde t}''(r)$ denote the corrections coming from ${\cal L}_{(4,2)}^-$, which can be of order ${\cal O}({\widehat m}^2)$, ${\cal O}({\widehat m} m_s)$, and ${\cal O}(m_s^2)$. In the standard case, the two last terms in Eq.~\rf{ggeta} would appear only at higher orders, and one may thus proceed as described in Ref.~\cite{bachir95}. At the order we are working, the quark mass ratio is then given as $r_{\mbox{\scriptsize st}} = r_2\equiv 2M_K^2/M_\pi^2-1\sim 25.9$~\cite{GL2}. Using the numerical values given in \cite{bachir95}, one obtains, from Eqs.~\rf{bachir1} and \rf{bachir2}, respectively, \vs{2}\begin{equation} {\widehat m} A_{4,{\mbox{\scriptsize st}}}-M_\pi^2(A_2-2A_3)_{\mbox{\scriptsize st}} \,=\, (-5.9\pm 1.8)\times 10^{-4} \\[2mm]\end{equation} and \vs{2}\begin{equation} 3{\widehat m} A_{6,{\mbox{\scriptsize st}}}+M_\pi^2(A_2-2A_3)_{\mbox{\scriptsize st}} \,=\, (-2.1\pm 0.4)\times 10^{-3}\ . \\[2mm]\end{equation} Upon using the experimental rate for $\eta\to\gamma\gamma$~\cite{PDG}, we then determine the combination \vs{2}\begin{equation} M_\pi^2(A_2-2A_3)_{\mbox{\scriptsize st}} \,=\, (1\pm 1) \times 10^{-3} .\lbl{A2minus2A3std} \\[2mm]\end{equation} Adding errors in quadrature, the previous results give \vs{2}\begin{equation} {\widehat m} t_{\mbox{\scriptsize st}} + M_\pi^2t'_{\mbox{\scriptsize st}} = (1.7 \pm 8.2)\times 10^{-3}\ , \\[2mm]\end{equation} which is four times more accurate than the value obtained directly from the experimental rate of $\pi^0\to\gamma\gamma$. For the separate pieces, we obtain \vs{2}\begin{equation} {\widehat m} t_{\mbox{\scriptsize st}} \,=\, (4.4\pm 10.8)\times 10^{-3}\, , \ \ M_\pi^2t'_{\mbox{\scriptsize st}} \,=\,(-2.7\pm 2.7)\times 10^{-3} \ .\lbl{stdresult} \\[2mm]\end{equation} In the generalized case, the analysis may, unfortunately, not be pursued quite that far. The main drawback are the corrections from ${\cal L}_{(4,2)}^-$, which in particular produce potentially large ${\cal O}(m_s^2)$ corrections to the $\eta\to\gamma\gamma$ decay amplitude, and on which the sum--rules \rf{bachir1} and \rf{bachir2} give no information. If one restricts the analysis to the order ${\cal O}({\mbox p}^5)$ precision, then the contributions from ${\cal L}_{(6,0)}^-$ are also absent, and the situation becomes even simpler than in the standard case, since ({\it cf.} Eq.~\rf{gencount}) the left--hand sides of the sum--rules \rf{bachir1} and \rf{bachir2} now only involve $A_4$ and $A_6$, respectively. Keeping in mind that $r$ is now a free parameter ($\alpha$ and $r$ are however related, see \cite{gchpt}), and taking the necessary inputs from \cite{bachir95}, the ${\cal O}({\mbox p}^5)$ determination of $A_4$ and $A_6$ reads \vs{2}\begin{equation} {\widehat m} A_4\,=\,-\frac{N_c}{32}\left( \frac{\vert\lambda (r) M_S^2 - (1-\lambda (r))^2M_\pi^2\vert}{M_P^2-M_S^2} \right)^{\frac{1}{2}}\frac{M_\pi}{M_P}\frac{{\cal A}(\pi'\to\gamma\gamma)}{{\cal A}(\pi\to\gamma\gamma)}\,+\,\cdots\ ,\lbl{A4gen} \\[2mm]\end{equation} \vs{2}\begin{equation} {\widehat m} A_6\,=\,-\frac{3\pi^2}{32}\frac{M_\pi}{M_{\eta '}}\left( \lambda (r) - \frac{\Delta_{\mbox{\scriptsize GMO}}}{(r-1)^2}\right)^{\frac{1}{2}}F_\pi {\cal A}(\eta'\to\gamma\gamma)\,+\,\cdots\ ,\lbl{A6gen} \\[2mm]\end{equation} where the ellipses stand for higher order corrections, $\lambda (r) = 2(r_2-r)/(r^2-1)$ and $\Delta_{\mbox{\scriptsize GMO}} \equiv (3M_\eta^2-4M_K^2+M_\pi^2)/M_\pi^2\sim -3.6$. For $r=r_2$, the above expressions give values compatible with the previous SChPT analysis. Furthermore, as $r$ decreases ({\it i.e} as $\alpha$ increases), the variation of ${\widehat m} t$ stays within the bounds given in Eq.~\rf{stdresult}. On the other hand, in the extreme case of a vanishing condensate, the ${\cal L}_{(6,0)}^-$ contributions to the two sum--rules also disappear, and the expressions \rf{A4gen} and \rf{A6gen} become exact at order ${\cal O}({\mbox p}^6)$. Finally, if we use the two sum--rules in order to express $A_4$ and $A_6$ in terms of $A_2-2A_3$ in the expression \rf{ggeta}, we obtain an estimate of a combination of $A_2-A_3$, ${\widehat m}^2{\widetilde t}''(r)$ and ${\widehat m}^2a_3$, which is not very sensitive to the value of $r$ and compatible with the value \rf{A2minus2A3std} obtained in the standard case for $M_\pi^2(A_2-2A_3)$. Thus, within reasonable error bars, the values of $t$ and $t'$ can be taken independent of $r$. For the numerical analyses presented in the text, we have used \vs{2}\begin{equation} {\widehat m} t\,=\, (6\pm 12)\times 10^{-3}\, , \ \ M_\pi^2t' \,=\,(-3\pm 3)\times 10^{-3} \ .\lbl{genresult} \\[2mm]\end{equation}
\section{Introduction} The measurement of proton momentum distributions by neutron scattering is analogous to the measurement of electron momentum distributions by Compton scattering\cite{pss} and measurement of nucleon momenta by Deep Inelastic Scattering \cite{is} and is known as Neutron Compton Scattering (NCS) or Deep Inelastic Neutron Scattering (DINS). All three techniques rely upon the fact that if the momentum transferred from the incident to target particle is sufficiently large, the impulse approximation (IA) can be used to interpret the data. In the IA, momentum and kinetic energy are conserved. From a measurement of the momentum and energy change of the neutron, the momentum of the target nucleus before the collision can be determined. DINS measurements have only become practical since the construction of intense accelerator based neutron sources, which have allowed inelastic neutron scattering measurements with energy transfers in the eV region\cite{hmt}. Energy transfers much greater than the maximum vibrational frequency of the target atom are required before the IA can be used to reliably determine the momentum distribution . At lower energy transfers the IA is no longer valid and is not related in a simple way to the observed scattering intensities. There have been a few pioneering studies on anisotropic systems at eV energy transfers\cite{rw,ph,isns,pal,ftm} but the analysis has been limited to fitting Gaussians to the observed data, or more generally fitting the data with model containing a few parameters, as was done for measurements on molecular hydrogen\cite{jm1}. We show here that an entire spectrum of anharmonic coefficients can be measured without recourse to any model, in addition to the widths of an anisotropic gaussian, thus describing an arbitrary anisotropic and anharmonic momentum distribution, The possibility of doing this for isotropic systems was first suggested by Reiter and Silver,\cite{rs} That possibility, for more general systems, is now a reality. We demonstrate this by measuring the momentum distribution for $KHC_2O_4$ where we obtain 14 anharmonic coefficients whose size varies by nearly two orders of magnitude, with at least 2-3$\sigma$ confidence levels for all but one. The experimental instrument is the EVS spectrometer at ISIS. The work presented here by no means represents the limits of resolution of the instrument, but rather the first experiments of this kind. Upgrades are planned in the near future that will significantly increase flux and counting efficiency. \section*{Theory of Measurement} The theoretical basis of neutron Compton scattering is the impulse approximation (IA), which is exact when the momentum transfer and energy transfer are infinite\cite{n,s,jm2}. The neutron scattering function S($\vec{q},\omega$), is related to the momentum distribution n($\vec{p}$) in the impulse approximation limit by the relation \begin{equation} S(\vec{q},\omega) = {M\over q}\int n(\vec{p})\delta(y-\vec{p}.\hat q)d\vec{p} = {M\over q} J(\hat {q}, y) \label{sqw} \end{equation} where y=${M\over q}(\omega-{q^2\over 2M})$, M is the mass of the target particle,q=$|\vec{q}|$, and $\hat{q}$=$\vec{q}/q$. DINS measurements on protons have a particularly simple interpretation, as the interaction of protons with other atoms can usually be accurately accounted for\cite{l,hrp,wls} in terms of a single particle potential and hence by a proton wave function.\cite{temp} From elementary quantum mechanics, n($\vec{p}$) is related to the Fourier transform of the proton wave function via, \begin{equation} n(\vec{p})={1\over(2\pi)^3)}|\int|\Psi(r)exp(i\vec p\cdot\vec r)d\vec r|^2 \label{npsi} \end{equation} and a DINS measurement of n($\vec{p}$) can be used to determine the wave function in an analogous way to the determination of real space structure from a diffraction pattern. If n($\vec{p}$) is known, and if the proton is in a site with reflection symmetry, so that the wave function can be assumed real, then in principle both the proton wave function and the exact form of the potential energy well in which the proton sits can be directly reconstructed.\cite{rs} With an asymmetric site such as potassium binoxalate, the phase information that is lost by taking the absolute value of the momentum wavefunction is irrecoverable, and we will not be able to reconstruct the potential directly. The n($\vec{p}$) obtained can, of course, be used to check any model potential. While the original formulation of the inversion problem\cite{rs} is complete as it stands, it is useful for the systems we will be dealing with to take into account the anisotropy of the system explicitly. The fundamental result that allows for a simple inversion of the Radon transform, J($\hat q, y$) to obtain n($\vec{p}$) makes use of a basis of Hermite polynomials and spherical harmonics in which the transform is diagonal. That is, a single term in the series for J($\hat q, y$) corresponds to a single term in the expansion of n($\vec{p}$). If we express J($\hat q, y$) in this basis as \begin{equation} J(\hat q, y) = {e^{-y^2}\over \pi^{1\over 2}} \sum\limits_{n,l,m} a_{n,l,m} H_{2n+l}(y)Y_{lm}(\hat q) \label{exp} \end{equation} then n($\vec{p}$) is given in the related basis of Laguerre polynomials as \begin{equation} \label{inv} n(\vec{p}) = {e^{-p^2}\over \pi^{3\over 2}}\sum\limits_{n,l,m}2^{2n+l}n!(-1)^n a_{n,l,m} p^lL_n^{l+{1\over 2}}(p^2) Y_{lm}(\hat p) \end{equation} where $\hat p$ and $\hat q$ are unit vectors. Clearly, since the expansions are complete, a distribution of the form \begin{equation} \label{anh} n(\vec p) = \prod\limits_i {e^{-p_i^2\over 2\sigma_i^2}\over (2\pi\sigma_i)^{1\over 2}} R(\vec p) \end{equation} with the $\sigma_i$ significantly different from each other, could be expanded in this form, but even if R($\vec p$) were 1, it would require a large number of terms in the series. To avoid this, we show that the anisotropy may be taken into account by a change of variables, so that the coefficients $a_{n,l,m}$ represent genuinely anharmonic contributions. Introducing the new variables \begin{equation} p_i^\prime = p_i /\sqrt2\sigma_i \end{equation} with n($\vec p$) defined as in Eq. (\ref{anh}) , defining $R^\prime(\vec {p^\prime}$)=R($\vec p(\vec {p^\prime}$)) and \begin{equation} \label{trans1} n^\prime(\vec p^\prime)={e^{-{p^\prime}^2}\over \pi^{3\over 2}}R^\prime(\vec p^\prime) \end{equation} we have \begin{equation} \label{trans} J(\vec{q}, y) = \int n^\prime(\vec {p^\prime}) \delta(y-\vec {p^\prime}.\vec {q^\prime})d\vec {p^\prime} \end{equation} where ${{q^\prime}_i} = q_i\sqrt2\sigma_i$. The right hand side of Eq. (\ref{trans}) is no longer a Radon transform, since $\vec q^\prime$ is not a unit vector. However, defining ${y^\prime} = y/\vert\vec q^\prime|$ we obtain \begin{equation} \label{fin2} J(\vec{q}, y) = {1\over \vert\vec q^\prime\vert}\int n^\prime(\vec {p^\prime}) \delta(y^\prime-\vec {p^\prime}.\hat {q^\prime})d\vec {p^\prime} = {1\over \vert\vec {q^\prime}\vert}J^\prime(\hat {q^\prime}, y^\prime) \end{equation} where $J^\prime(\hat {q^\prime}, y^\prime)$ is the Radon transform of the isotropic(in it's gaussian component) but anharmonic distribution $n^\prime(\vec {p^\prime})$. If $ {\hat q}$ is specified as a unit vector in the usual spherical coordinates, then \begin{equation} \label{fin} \vert\vec q^\prime\vert = \sqrt2\big((\sigma_1 sin(\theta)cos(\phi))^2+(\sigma_2 sin(\theta)sin(\phi))^2+(\sigma_3 cos(\theta))^2\big)^{1\over 2} \end{equation} Our procedure is to expand $J^\prime(\hat {q^\prime}, y^\prime)$ in hermite polynomials, as in Eq. (\ref{exp}), and least squares fit the data, S($\vec{q},\omega$), using Eqs. (\ref{sqw},\ref{fin2},\ref{fin}), to obtain the parameters, $\sigma_i, a_{n,l,m}$. $n^\prime(\vec {p^\prime})$ can then be reconstructed as in Eq. (\ref{inv}), and we thus obtain n($\vec{p}$) as in Eq.(\ref{anh}) with R($\vec p$)= $R^\prime(\vec {p^\prime}(\vec p))$. That this is a practical procedure will be demonstrated below. \section{Measurements} The measurements were performed on the electron volt spectrometer eVS\cite{me} at the ISIS neutron source. On EVS the energy of the scattered neutron is fixed by a resonance filter difference technique\cite{stb}. The final neutron velocity and energy are related by $E_1=m{\nu_1}^2/2$ where m is the neutron mass. The energy of the incident neutron is determined from a measurement of the neutron time of flight via the equation \begin{equation} \label{tof} t={L_0\over\nu_0}+{L_1\over\nu_1} \end{equation} where t is the measured time of flight $,L_0$ and $L_1$ are the lengths of the incident and the scattered flight paths of the neutron, $\nu_0$ and $\nu_1$ are the speeds of the incident and scattered neutrons. Then \begin{equation} \label{endef} \omega=m({\nu_0}^2-{\nu_1}^2)/2 \end{equation} and \begin{equation} \label{qdef} q=m({\nu_0}^2+{\nu_1}^2+2\nu_1\nu_0cos\theta)^{1\over 2} \end{equation} where $\theta$ is the scattering angle. From these equations $\omega$ and $\vec q$ can be determined for a given time of flight t , if the instrumental parameters $L_0,L_1,\theta$ and $E_1$ are known. Hence from the count rate at a given time t, J($\hat q, y$) can be determined. On eVS the detectors are situated in the horizontal plane and hence $\vec q$ is always horizontal. By orienting the sample with a chosen crystal axis vertical, it is possible to measure J($\hat q, y$) for $\vec q$ in whichever plane, relative to the sample, one chooses. A time of flight scan at a particular angle for a given detector does not correspond, however, to a particular direction of $\vec q$. There is significant curvature of this scan through the proton momentum space since the direction of $\vec q$ varies significantly over the data region. Time of flight spectra for eight adjacent detectors at angles between 35 and 55 degrees scan through the atomic momentum momentum space of the proton as illustrated in figure 1 \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=gr1.ps,width=100mm} \caption{Scan pattern in momentum space for detectors at a fixed angle as the time of flight is varied } \label{fig1} \end{figure} A complete scan over the proton momentum space is constructed by combining a number of data sets, taken with the sample rotated about the vertical axis by appropriately chosen angles. The reported measurements were made using two banks of 8 $ Li^6$ doped glass scintillator detectors which were symmetrically placed on each side of the incident beam at scattering angles between $35^o$ and $55^o$. For DINS studies of protons it is necessary to site the detectors at forward scattering angles since the hydrogen scattering cross section is strongly anisotropic at eV incident energies, with virtually no back scattering. This restriction is a kinematic consequence of the closeness of the mass of the neutron and the hydrogen atom and does not apply to heavier atoms. The resolution function of the instrument is determined by the uncertainties in the measured values of the time of flight t and the distribution of $L_0,L_1,\theta$ and $E_1$ values allowed by the instrumental geometry and analyser foil resolution. Uncertainties in $L_0$ arise primarily from the finite depth of the neutron moderator, those in $L_1$ and $\theta$ from the finite sample and detector sizes and those in t from jitter in the detector electronics. All resolution components can be determined by calibration measurements and all except the energy component can be approximated by Gaussians, without significant error. A 0.015 mm thick gold foil provided a Lorentzian energy resolution function at $E_1=4908meV$ , with a peak HWHM of 136 meV. The Gaussian and Lorentzian resolution components in momentum space y ,are listed in table 1 for two angles representative of the range of angles employed. The resolution is dominated by the energy component which varies strongly with scattering angle. The second most important contribution comes from the angular resolution of the spectrometer and is independent of angle. The momentum and energy transfers at the centre of the hydrogen response peak are also listed for the different angles. \vskip .25in {\bf Table 1.} The resolution widths are the Lorentzian HWHM for (RL) and the Gaussian standard deviation for other parameters (RG) . The momentum q and energy transfer ( at the scattering angles $35^o$ and $55^o$ are also given.) \begin{center} \begin{tabular}{||c|c|c|c|c||} \hline \hline Angle & $R_G(\AA^{-1})$ & $R_L(\AA^{-1})$ & q $(\AA^{-1})$ & $ \omega$ (eV)\\ \hline \hline $ 35^0$ & 0.61 & 1.08 & 34.1 & 2.41\\ \hline $ 55^0$ & 0.55 & 0.55 & 48.8 & 4.92\\ \hline \hline \end{tabular} \end{center} The raw data contains signals from all the atoms in the scattering sample and from the cryostat background. Fortunately the energy transfer to hydrogen is much greater than that to other atomic masses and the proton signal is well separated from that due to other masses. The contribution from all components other than hydrogen is subtracted by fitting a sum of Gaussians convoluted with the instrument resolution function to the data and subtracting off the fitted contribution to other peaks. There is also a small contribution to the data from a second gold resonance at 60 eV which can be seen at ~100 $\mu$sec and this is also fitted and subtracted from the data. \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=gr2.ps,width=100mm} \caption{ The sum of data from 8 detectors at scattering angles between $38^o$ and $55^o$ is shown as the dotted line. The data after subtraction of the contribution from atoms with higher masses and the 60 eV resonance data is shown as the full line.The total data set for a single plane consisted of 36 such spectra. The FWHM of the Lorentzian resolution function is ~6 microsec.} \label{fig2} \end{figure} The data for each scan was converted into a distribution in the momentum space of the crystal as described above. Complete coverage of the plane was achieved by combining six runs which were taken at steps of $23^o$ sample rotation. A contour plot of the data derived from the six runs is shown in figure 3. This was produced by binning the counts from different detectors and sample orientations in the appropriate pixel of the momentum space \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=gr3.ps,width=100mm} \caption{Contour plot of data in one plane of potassium binoxalate. The apparent double peak near the origin is an artifact of the method of plotting the data } \label{fig3} \end{figure} The data has been corrected for sample attenuation, but still contains errors due to small deviations from the Impulse Approximation which are present at the finite momentum transfers of the measurement. These tend to introduce small asymmetries into the data set at the ~1-5\% level, thereby removing the exact inversion symmetry of the Compton profile. It has been shown by Sears\cite{s2} that most of these effects are removed by symmetrisation of the data about the origin. This procedure cannot be followed for our data sets as the scan in a single detector is curved and different points in the crystal plane which are related by inversion symmetry may have been measured in different detectors, with different experimental resolution. However by fitting to a J($\hat q, y$ with inversion symmetry, as discussed below, we automatically include a correction for deviations from the impulse approximation. Any asymmetries are ignored by the fit and should not affect the values of the fitted parameters. \section{Fitting Procedures} Eqs.(\ref{exp},\ref{inv}) hold quite generally for Radon transform pairs, but physical requirements in the present context restrict the allowed coefficients. Since $J(\vec{q}, y)$ is an even function of y, l is restricted to even values, and since $J(\vec{q}, y)$is real, the $a_{n,l,m}$ for $\pm m$ must be equal. In fact, if there are residual final state effects in the data, the data will not be symmetric. Restricting the coefficients in this way will therefore eliminate these residual effects. Lifting the restrictions and fitting the data allows one to measure the extent of these effects.For potassium binoxalate, data was taken for three perpendicular planes oriented parallel to the crystal axes The procedure followed was to perform a simultaneous fit to the 6x16x3=288 separate time of flight spectra to an expansion of the form given in Eq.(\ref{exp}), convoluted with the instrument resolution function The actual fitting procedure requires that three of the parameters $\sigma_i, a_{n,l,m}$ be fixed, as they are not all independent. That is, if one varies the $\sigma_i$ abitrarily, there will always be a set of $a_{n,l,m}$ that will fit the data, the particular values that fit depending on the choice of the $\sigma_i$. One could obtain fixed values for the $\sigma_i$ by first fitting the data to a gaussian, i.e. $R^\prime(\vec p^\prime)=1$, and then varying only the $a_{n,l,m}$, in which case all such terms would in principle be needed. We find in practice, that better results are obtained, i.e. fewer anharmonic coefficients needed and more rapid convergence of the series for $R^\prime(\vec p^\prime)$, when the $\sigma_i$ are allowed to vary but the first three anharmonic coefficients, $a_{1,0,0}, a_{0,2,0}, a_{0,2,2}$ are set to zero. In fact, these coefficients are not really anharmonic coefficients at all. They could always be eliminated by a shift of the $\sigma_i$ and an adjustment of higher order coefficients. This procedure thus has the virtue as well of producing only genuinely anharmonic corrections to a gaussian fit. The datasets as they are presently obtained in the EVS spectrometer at ISIS, are obtained one plane at a time(see discussion of experimental apparatus). That is $\vec q$ varies within a plane, and a range of y values is taken such that $J(\vec{q}, y) $is negligible outside this range. There is a very high density of points, which for the purposes of the present discussion we can take to be continuous. The question then arises, 'How many planes of data are needed to determine a specified number of coefficients, and at what angles to each other should they be?' We can see the point of the question by considering first a single plane and looking at the fit to the leading anharmonic coefficient. From Eq.(\ref{exp}) we see that there are six independent coefficients multiplying $H_4$, i.e. the coefficients of $(Y_{00},Y_{20}, Y_{22}, Y_{40}, Y_{42}, Y_{44})$. Let us say that our coordinate system is chosen so that the plane measured is treated as an xz plane. Then in terms of the variable $\theta$ there are only three independent fourier coefficients that can be present in the data for the coefficients of $H_4$, i.e. the coefficients of (1, $cos(2\theta), cos(4\theta)$ for instance. Therefore three of the coefficients are not independent. The complete set of coefficients cannot be determined by the data. This is of course due to the fact that there is no information as to the behaviour of $J(\vec{q}, y)$ for nonzero azimuthal angles $\phi$ in the data, so we shouldn't expect the fitting procedure to provide it. The data provides a complete description of $J(\vec{q}, y)$ only if this function is rotationally invariant about the z axis. Actually, what is required is only that $J^\prime(\hat {q^\prime}, y^\prime)$ be rotationally invariant, since the fit is done in the primed coordinate system. If this is the case, then all coefficients of $Y_{lm}$ with non-zero values of m must be zero. We see that there are only three remaining possibly non zero coefficents, which can all be determined. For higher order terms as well, keeping only the coefficients with m=0 provides all the independent terms needed to fit the data, and the resulting fit, of course, is rotationally symmetric about the z axis. If the data is not known to be rotationally symmetric, additional planes of data must be taken to determine even these lowest order coefficients. In general, whenever we take another plane of data, we might expect to obtain 3 more independent measurements of the coefficients of $H_4$, 4 independent measurements of the coefficients of $H_6$, and in general, k+1 measurements of the coefficients of $H_{2k}$. (k+1 being the number of independent fourier components in the data for that value of k). Since the number of $a_{n,l,m}$ that are to be determined for 2n+l=2k is (k+2)(k+1)/2, it appears that (k+2)/2 planes are needed to measure all coefficients up to $H_{2k}$. The angles between the planes must be chosen, however, so that the measurements are really independent. For instance, if k=2, it would appear that two planes would suffice, but if they are chosen as the xz and yz planes, they do not provide independent measurements of the coefficients. This may be seen by observing that the sum of the data from the two planes gives three independent fourier coefficients to determine four independent $a_{n,l,m}$, the coefficients of $(Y_{00},Y_{20}, Y_{40}, Y_{44})$, since the coefficients of $Y_{22}$ and $Y_{42}$ cannot affect this sum. The difference of the data on the two planes gives three equations for the two coefficients of $Y_{22}$ and $Y_{42}$. A better choice for the planes would be $\phi=0$ and $\phi=\pi/4$, which would allow the determination of all the coefficients. If there is some symmetry in the problem, one may be able to use perpendicular planes if the symmetry axis is chosen appropriately with respect to the common axis of the two scattering planes. For instance, if there is tetragonal symmetry present, and the symmetry axis is chosen perpendicular to the common axis, one can obtain all the allowed coefficients up to k=4. One can show that three perpendicular planes do in fact suffice to determine the coefficients up to k=4 in the general case, without any symmetry to reduce the number of allowed coefficients. The question of whether this is an optimum configuration of planes or not, we will leave to another time. Including the three $\sigma_i$, a three plane measurement allows 34 coefficients to be measured. \section{Measurement Errors} The uncertainty in the measurement of n($\vec p$) at some point $\vec p$ is due to the uncertainty in the measured coefficients. Denoting an arbitrary coefficient by $\rho_i$, we have \begin{equation} \label{uncn} \delta n(\vec{p})= \sum\limits_{i} {\delta n(\vec{p})\over \delta\rho_i} \delta\rho_i \end{equation} The fitting program, after a minimum is obtained with some set of coefficients, calculates the correlation matrix $<\delta\rho_i\delta\rho_j>$ by varying the coefficients slightly and calculating the curvature of the chi-square of the fit. \cite{dev}. Hence, the variance in the momentum distribution is \begin{equation} \label{deln} <\delta n(\vec{p})^2>= \sum\limits_{i,j} {\delta n(\vec{p})\over \delta\rho_i} {\delta n(\vec{p})\over \delta\rho_j}<\delta\rho_i\delta\rho_j> \end{equation} There are of course, potential systematic errors that could enter the measurement, such as multiple scattering effects, or an error in determining the resolution function. The former must be handled with good experimental design, and if small, can be corrected for. We have done measurements on samples whose thickness differed by a factor of two, with no significant differences in the observed scattering. Multiple scattering effects would also lead to asymmetries in J($\hat q, y$) that are not observed. The resolution function has been studied extensively, and is believed to be known accurately.In the present measurement, the resolution width is between 15 and 25\% of the width of the distribution measured. A further source of error not contained in the estimate in Eq.(\ref{deln}) is the truncation of the series used to fit the data to include only terms up to 2n+l=8. To get an idea of the seriousness of this, and to test the fitting procedures and software, we have generated synthetic data from a known momentum distribution that corresponds to an asymmetric double well, and convolved it with the instrumental resolution function. The data was then analyzed by the means described above, and the extracted n($\vec{p}$) compared with the input. The input n($\vec{p}$) corresponded to a spatial wave function consisting of two displaced gaussians with the same variance and a relative weight r=.5. The explicit form is \begin{equation} \label{dpotn} n(p_x,p_y,p_z)={(1+r^2+2rcos(2p_za))\over{ (1+r^2+2re^{-{a^2\over 2{\sigma_z}^2}}})} \prod\limits_i {e^{-{p_i^2\over 2\sigma_i^2}}\over (2\pi\sigma_i)^{1\over 2}} \end{equation} where $\sigma_x=4.6, \sigma_y=4.0,\sigma_z=6.0 $ and a=.15 in units of inverse angstroms and angstroms, respectively. This form is rather similar to the actual form of the data we will analyze. The coordinate system z axis was chosen to be identical with the crystal z axis. The comparison is shown in fig 4. The extracted n($\vec{p}$) is plotted with a dotted line, the input n($\vec{p}$) above with a dashed line. As may be seen from the figure, there is essentially no error due to the truncation of the the expansion. Of course, an input with more variation might require including higher terms in the expansion, and hence taking more planes of data. The actual data we have obtained for potassium binoxalate has less variation than the simulation and the truncation error will be negligible. In the other directions, the fit is rigorously gaussian, since all coefficients with $m\not=0$ were zero. We note that the measured $\sigma_z$ parameter was 4.47, not 6.0. It is simply a fitting parameter. The overall momentum distribution has physical significance, the individual parameters may not. In fact, we have gotten the same degree of fit with the crystal z axis aligned along the coordinate y axis. In this case, all the coefficients are non-zero, but the resultant n($\vec{p}$) is identical to the one displayed above. \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=simplotz.ps,width=100mm,angle=-90} \caption{Comparison of input n($\vec{p}$), given by Eq.(16)with reconstruction of n($\vec{p}$ )using fitting procedure. The z coordinate axis is chosen to be the double well axis.} \label{fig4} \end{figure} A final source of systematic error involves the possibility of finding a false minimum with the fitting procedure. With such a large number of parameters, there is the possibility that the program will home in on a local minimum and miss the true minimum. We are using straightforward gradient methods for the search, which would have difficulty with a very rugged chi-squared landscape. We do not appear to have such a landscape for these problems, and certainly not for the large parameters, such as the $\sigma_i$, or the largest anharmonic coefficients, whose values appear to be quite robust with regards to different paths to the minimum. One can check for this problem by orienting the fitting coordinate system differently with respect to the crystal axes, as was done for the test data above. In this case, all the coefficients will be different, but the final n($\vec{p})$ must be the same. There is also the fact that n($\vec{p})$ must be positive, and a spurious fit that leads to significant negative values can be rejected. We can also eliminate some parameters which are consistently much smaller than their variance,and which can be set to zero without affecting significantly the chi-square, thus reducing the dimensionality of the space. We believe there are no problems of this sort with the fits we will present. \section{Results for Potassium Binoxalate} Potassium Binoxalate, $KHC_2O_4$, is a hydrogen bonded system in which the hydrogen sits in an asymmetric position between two oxygen atoms.\cite{emd}.The crystal is monoclinic. We will choose a primitive cell for which the bond axis, that is the line joining the two oxygen atoms, is essentially aligned with the c axis of the crystal.\cite{cell} We choose this axis for the z axis of our coordinate system. Three planes of data were taken at right angles to each other, with 69,632 data points in all. One of these is the bc plane, where the b axis is the unique axis, and the other two are the $a^*b$ and $a^*c$ plane, where the $a^*$ axis is perpendicular to the bc plane. These were fit with the methods described above to give the momentum distribution. The measurements were done at 10 deg K and hence there are no significant finite temperature corrections to the ground state momentum distribution due to excited states. Note that the coordinate system used to describe the momentum distribution need not have the symmetry of the crystal. The local symmetry of the site is only a two-fold rotation. We show the values of the fitted coefficients in tables 2 and 3, along with the rms uncertainty in their values. This is included only to give some sense of the individual parameter. The error in n($\vec{p}$)is given by Eq.(\ref{uncn}) and includes the effect of the correlations between coefficients, whereas the figures cited in the table are only the diagonal correlation coefficients. It can be seen, that many of the measured coefficients have been determined at the 2-3 $\sigma$ level of confidence, some at higher levels, and only one at a 1$\sigma$ level. Coefficients that are set to zero in the fitting procedure were found to have values smaller than their variance by at least a factor of 2, and setting them to zero did not significantly change the minimum value of chi-square. The goodness of fit to a sample of the data is shown in Fig. 5. \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=gr5.ps,width=100mm,angle=-90} \caption{Data fitted by method described above. The data is a composite of data points in a 10deg wedge about the z axis. } \label{fig5} \end{figure} We have compared the fitted prediction, convolved with the instrumental resolution, and the data, for the cumulative sum of data in a 10 degree wedge along the hydrogen bond direction. The measured momentum distribution along the axes of the measurement planes, are shown in Figs. 6-9, and contour plots along coordinate planes in Figs. 10-13. \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=pgz.ps,width=100mm,angle=-90} \caption{Momentum distribution for potassium binoxalate along the axis shown. The momentum is in units of $\AA^-1$, n($\vec{p}$ is in arbitrary units. The errors are calculated as in Eq.(15) for the parameters that are significant in Tables 2 and 3. The lower curve is the anharmonic component. } \label{fig6} \end{figure} \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=pgx.ps,width=100mm,angle=-90} \caption{Momentum distribution for potassium binoxalate along the axis shown. The momentum is in units of $\AA^-1$, n($\vec{p}$) is in arbitrary units. The errors are calculated as in Eq.(15) for the parameters that are significant in Tables 2 and 3. The lower curve is the anharmonic component. } \label{fig7} \end{figure} \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=pgy.ps,width=100mm,angle=-90} \caption{Momentum distribution for potassium binoxalate along the axis shown. The momentum is in units of $\AA^{-1}$, n($\vec{p}$) is in arbitrary units. The errors are calculated as in Eq.(15) for the parameters that are significant in Tables 2 and 3 }. The lower curve is the anharmonic contribution. \label{fig8} \end{figure} \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=pg2xz.ps,width=80mm,angle=-90} \caption{Momentum distribution for potassium binoxalate along the axes shown. The momentum is in units of $\AA^{-1}$, n($\vec{p}$) is in arbitrary units. } \label{fig9} \end{figure} \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=pg2yz.ps,width=80mm,angle=-90} \caption{Momentum distribution for potassium binoxalate along the axes shown. The momentum is in units of $\AA^{-1}$ } \label{fig10} \end{figure} \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=pg2xy.ps,width=80mm,angle=-90} \caption{Momentum distribution for potassium binoxalate along the axes shown. The momentum is in units of $\AA^{-1}$. } \label{fig11} \end{figure} \newpage \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=pgcxy.ps,width=80mm,angle=-90} \caption{Anharmonic contribution to the momentum distribution for potassium binoxalate along the axes shown. The momentum is in units of $\AA^{-1}$. } \label{fig12} \end{figure} \begin{figure}[h] \hspace{.2 \hsize} \psfig{figure=pgcxz.ps,width=80mm,angle=-90} \caption{Anharmonic contribution to the momentum distribution for potassium binoxalate along the axes shown. The momentum is in units of $\AA^{-1}$. } \label{fig13} \end{figure} \newpage The coefficients that were measured are given in the tables below. \vskip .25in {\bf Table 2.} The harmonic fitting coefficients and the variances in their values as measured \begin{center} \begin{tabular}{||c|p{.6in}|p{.6in}||} \hline \multicolumn{3}{||c||}{\bf Harmonic Coefficients}\\ \hline i&$\sigma_i$&$\delta\sigma_i$\\ \hline x&4.509&.0235\\ y&4.869&.0254\\ z&5.351&.0449\\ \hline\hline \end{tabular} \end{center} \smallskip \vspace{.5in} {\bf Table 3.} The anharmonic fitting coefficients and their variances as measured. \begin{center} \begin{tabular}{||c|c|c|p{.6in}|p{.6in}||} \hline \multicolumn{5}{||c||}{\bf Anharmonic Coefficients}\\ \hline n & l & m & $a^{\prime}_{n,l,m}$ & $\delta a^{\prime}_{n,l,m}$\\ \hline 2 & 0 & 0 & 0.0000 &0 .0000\\ 1 & 2 & 0 & -0.1443&0.0377\\ 1&2&2&0.0000 &0.0000 \\ 0&4&0&0.0510 &0.0086 \\ 0&4&2& -0.0564&0.0095 \\ 0&4&4&0.0000 &0.0000 \\ 3&0&0&-0.1356 &0.0098 \\ 2&2&0&-0.0296 &0.0105 \\ 2&2&2&0.0000 &0.0000 \\ 1&4&0&-0.000 &0.0000 \\ 1&4&2&0.0000 &0.0000 \\ 1&4&4&0.0000 &0.0000 \\ 0&6&0&-0.0184 &0.0065 \\ 0&6&2&-0.0127 &0.0028 \\ 0&6&4&-0.0388 &0.0129 \\ 0&6&6&0.0000 &0.0000 \\ 4&0&0&-0.0290 &0.0029 \\ 3&2&0&0.0000 &0.0000 \\ 3&2&2&0.0000 &0.0000 \\ 2&4&0&-0.0016 &0.0013 \\ 2&4&2&0.0069 &0.0013 \\ 2&4&4&0.0000 &0.0000 \\ 1&6&0&-0.0049 &0.0021 \\ 1&6&2&0.0000 &0.0000 \\ 1&6&4&-0.0098 &0.0036 \\ 1&6&6&0.0000 &0.0000 \\ 0&8&0&0.0000 &0.0000 \\ 0&8&2&0.0000 &0.0000 \\ 0&8&4&0.0000 &0.0000 \\ 0&8&6&0.0000 &0.0000 \\ 0&8&8&-0.0075 &0.0028 \\ \hline\hline \end{tabular} \end{center} \smallskip \vskip .1in Note that there are 14 anharmonic coefficients that are measureable, 13 of which are at the $2\sigma$ level at least, and that the harmonic coefficients are measured to better than 1\%. \section{Conclusion} We have shown how the method of analysis of DINS data suggested in \cite{rs} can be extended to anisotropic momentum distributions, and have applied this method to an analysis of the hydrogen bond in potassium binoxalate. The results demonstrate that DINS, as it is implemented now at ISIS, is capable of detailed, model independent, measurement of the momentum distribution for hydrogen, and by inference, other light atoms. These measurements required about four days of beam time, and are the first such measurements to be analyzed in this way. The count rates, resolution and detetector efficiency can be, and are scheduled to be, significantly improved. The data can be analyzed in less than a day. The DINS technique thus provides a practical means of accessing precise information about the anharmonicity of local potentials and can provide a check of any theoretical calculation of these potentials at a level of accuracy and detail that has not been possible previously. \section{Acknowledgements} We would like to thank Devinder Sivia for assistance in analyzing the data, Steve Bennington and John Tomkinson for useful discussions, and Professors R.G.Delaplane and H. Kuppers for providing the crystals.
\section{Introduction} The $Q$ counting scheme introduced just over one year ago by Kaplan, Savage and Wise (KSW) \cite{KSW} represents an important advance in the development of effective field theory techniques for nuclear physics. The approach is systematic, it builds in approximate chiral symmetry and chiral power counting, it solves Weinberg's \cite{Weinberg} ``unnatural scattering length problem'', and, in principle, it provides {\it a priori} estimates of errors for observables since one works to fixed order in $Q/\Lambda$. This last feature is extremely important since in Weinberg's initial formulation \cite{Weinberg} it was unclear how to make such error estimates. \footnote{Of course, the principal difference between $Q$ counting and the Weinberg scheme is that in Weinberg's approach the potential is iterated to all orders whereas in $Q$ counting only the leading term is iterated to all orders; all other operators are treated perturbatively. If $Q$ counting is valid, a scheme like Weinberg's if it can be systemat ically implemented will, at worst, add uncontrolled higher order contributions which do not spoil the systematic error estimates. Thus even if one is using Weinberg's approach, if one wants to make simple error estimates one may resort to $Q$ counting for the error estimate, provided the system is in a regime where it is valid. This view of error estimation in the Weinberg scheme is featured prominently in the discussions at this workshop.} To a considerable practical extent the major advantage to using EFT technology as opposed to unsystematic models is the ability to specify the accuracy of one's predictions. Thus, in this talk, I will focus entirely on $Q$ counting and not on the many beautifully accurate calculations based on Weinberg's approach implemented with a finite cutoff \cite{cutcalcs}. The scheme introduced by KSW may be divided up into two parts. The first is $Q$ counting and the second is a set of technical tricks to implement the $Q$ counting. These technical tricks are rather unusual; they are based on the so-called PDS scheme for doing subtractions in dimensional regularization. Apart from the very peculiar prescriptions required (subtracting the poles as $d=3$!) the formalism is not completely transparent in terms of the physics. Of course, provided the scheme is consistent and we are in the regime for which $Q$ counting is valid, one should get the same results for any scheme which implements the $Q$ counting. Two other approaches to $Q$ counting have been tried---the OS scheme with dimensional regularization of Mehan and Stewart\cite{MS} and a cutoff scheme in configuration space \cite{CH1}. All the schemes give the same results at fixed order in $Q$ counting. Thus, ultimately the physics turns on whether or not the $Q$ counting is working. The $Q$ counting scheme is straightforward: \begin{eqnarray} Q \, & \sim \, 1/a \nonumber \\ Q \, & \sim \, m_\pi \nonumber \\ Q \, & \sim \, p \label{Qcount} \end{eqnarray} where $a$ is the scattering length--either singlet or triplet, and p represents external momenta. For partisans of the OS scheme, you can just as well use $\gamma =\sqrt{M B}$ where $B$ is the magnitude of the binding energy, in place of $1/a$. All other scales are assumed to be heavy and will collectively be denoted as $\Lambda$. The expansion implied by $Q$ counting is in $Q/\Lambda$. Now the key point of $Q$ counting is that since $p$, $m_\pi$ and $1/a$ are all of the same order, at any order in $Q$ counting we have $ 1/(p a)$ and $p/m_\pi$ to all orders. One important feature of $Q$ counting is that while one needs to iterate the lowest order contact term to all orders to get a consistent result, all higher order contributions, including those from the pion can be treated perturbatively. $Q$ counting has been used to calculate a number of observables \cite{KSW,MS,Qpred} and generally seems to have real predictive power. At first sight this would appear to rule out the possibility that $m_\pi \sim \Lambda$. However, most ``vanilla'' observables principally test the $1/(a \Lambda)$ part of the theory. Clearly, it is important to identify observables which are principally sensitive to the $m_\pi/\Lambda$ parts of the theory and to rigorously test the chiral expansion. Recall that only the chiral part of the $Q$ counting is really understood in terms of QCD. In $Q$ counting, the small value of $1/a$ is treated as an essential fact of life that we cannot ignore. >From the QCD level, however, this fact of life is seen as essentially an accident. In contrast, the chiral physics is understood directly in terms of the small quark masses in the QCD Lagrangian along with spontaneous symmetry breaking. The central theme of this talk is that the effective range expansion---namely the expansion of $p \cot (\delta )$ as a power series in energy---is a good place to test whether the chiral part of $Q$ counting is under control. The effective range expansion (ERE) is a good place to look at pionic effects for a number of reasons. The expansion may be written: \begin{equation} p \cot (\delta ) \, = \, -\frac{1}{a} \, +\, \frac{1}{2} r_e \, p^2 \, + \,v_2 p^4 \, + \, v_3 p^6 \, + \, v_4 p^8 + \, \ldots \end{equation} Simple $Q$ counting shows that the scattering length term is order $Q^1$ while all other terms in the ERE are at least order $Q^2$. When one includes pions explicitly the same counting holds; all terms except the first are ${\cal O}(Q^2)$. It is important to note the distinction between the $Q$ expansion and the momentum expansion in the ERE. They differ precisely because the $Q$ expansion has $k/m_\pi$ and $1/(k a)$ to all orders while in the momentum expansion they are multiplied out. One immediately deduces that $v_n \sim Q^{-2 n + 2}$. Moreover all of the $v_i$ coefficients in the expansion diverge in the chiral limit of $m_\pi \rightarrow 0$. Hence they are pion dominated quantities and should provide a test of the chiral part of $Q$ counting. \section{Scales in Nuclear Physics} Before coming to the effective range expansion in this approach, it is useful to look a bit more closely at the various scales underlying $Q$ counting. Formally, there are two light scales intrinsic to the problem, $1/a$ and $m_\pi$. In $Q$ counting they are both formally of the same order namely ${\cal O}(Q)$. But emprically for both the triplet and singlet channel, \begin{equation} m_\pi \gg 1/a \end{equation} with $m_\pi a \approx 4$ for the triplet channel and $m_\pi a \approx 16$ for the singlet channel. This raises the following logical possibility; \begin{eqnarray} m_\pi \, & \sim & \, \Lambda \nonumber\\ 1/a \, & \ll & \, \Lambda \label{cond}\end{eqnarray} {\it i.e.}, that there is no scale separation between pionic scales and the ``short distance scales'' but there is a good scale separation between them and $1/a$. Of course, an immediate prejudice is that the first relation in eqs.~(\ref{cond}) must be wrong; chiral scales are intrinsically long compared to typical hadronic scales. However, what is relevant here for $\Lambda$ is not hadronic scales, but {\it nuclear} scales. The relationship of nuclear scales to QCD is quite obscure, but it is certainly true that typical nuclear mass scales are much lower than typical hadronic scales. If conditions in eqs.~(\ref{cond}) turn out be true one would expect that the parts of the theory which depend on $ 1/(a \Lambda)$ will work quite well, while the parts which depend on $m_\pi/\Lambda$ will converge slowly or not at all. Note, if it turns out that $ m_\pi \sim \Lambda $, there is nothing in principle wrong with the $Q$ counting formalism and PDS. It would simply not be useful for real world situations. Of course, one could play God and consider a world in which the pion is much lighter than in nature and then one would have real predictive power. In principle, if lattice technology improves, one could calculate properties in such an artificial world from first principles and could use the $Q$ counting technology to make predictions for this world. Before looking at explicit calculations, we should ask whether it is reasonable to suppose that $m_\pi \sim \Lambda$. Ultimately, this question comes down to whether $1/a$ and $m_\pi$ are the only light scales in nuclear physics. The answer appears to be ``no''. Numerically: $$\frac{1}{m_\pi} \approx 1.5 {\rm fm} \; \; a^s \approx -23 {\rm fm} \; \; a^t \approx 6 {\rm fm} $$ where the superscript s (t) refers to the singlet (triplet) channel. Compare these with the effective ranges: $$r_e^s \approx 2.7 {\rm fm} \; \; r_e^t \approx 1.6 {\rm fm} $$ It is apparent that $m_\pi r_e \sim 1$; if $r_e \sim 1/\Lambda$ there is a serious potential problem. Of course, it is possible that the large numerical size of $r_e$ is itself a reflection of chiral physics. For example, if $r_e \sim 1/m_\pi$ there would be no problem. One can use $Q$ counting itself to answer the question of how $r_e$ behaves. At leading nonvanishing order it is given by \cite{MS,CH1} \begin{equation} r_e \, = \, {\cal O}(\Lambda) \, + \, \frac{g_A^2 M}{4 \pi f_\pi^2 a^2 m_\pi^2} \, - \ \frac{g_A^2 M}{3 \pi f_\pi^2 a m_\pi} \, =\,{ \cal O}(Q^0) \end{equation} Note that although there {\it is} a chiral enhancement---the last two terms diverge in the chiral limit---it is compensated for by factors of the scattering length in the denominator. Thus, one expects in the context of $Q$ counting the effective range to be a short distance scale. In practice, however, it is larger than $1/m_\pi$. This suggests, but does not prove, that the chiral scale is not well separated from ``short distance'' scales. There is another way to see that ``short distance'' scales may be comparable to the pion mass scale. Consider the typical scales in so-called realistic N-N potential models, {\it i.e., those which fit the scattering data}. If you look at the non--one-pion-exchange part of the potential it is, in fact, larger than the OPEP potential for a distance less than $\sim 1.5-2$ fm. Since $1/m_\pi$ is comparable to, or shorter than, this distance we again appear to have evidence that non-chiral supposedly short distance scales are comparable to $1/M_\pi$. One might argue that the central potential contains two-pion-exchange physics (suitably reparameterized) in the potential model. However, if $Q$ counting is valid, that contribution is chirally suppressed. This argument appears to be model dependent as it is based on ``typical'' potential models. There is a model independent way to constrain the short distance physics.\cite{scal} Consider {\it any} nonrelativistic potential, including possible non-local potentials. Write the potential as the sum of an OPEP potential and some short distance potential with the constraint that the short distance potential vanishes beyond some distance $R$: \begin{eqnarray} V(\vec{r},\vec{r'}) \, = \, V_{\rm OPEP}(\vec{r}) \, \delta(\vec{r} - \vec{r'}) + V_{\rm short}(\vec{r},\vec{r'}) \nonumber \\ \nonumber \\ \; \; \; {\rm with} \; \; \; V_{\rm short}(\vec{r},\vec{r'}) = 0 \;\; \; {\rm for } \; \; \; r, r' \, > \, R. \end{eqnarray} Now suppose that this potential is inserted to a Schr\"odinger equation and used to solve for singlet phase shifts. A remarkable theorem can then be proved, namely that if the short distance potential fits the scattering length and effective range there is a minimum value for $R$. For real world values one can deduce the $R > $1.1 fm \cite{scal}. Moreover the derivation of this bound shows that it is unsaturatable so one expects $R$ to be significantly more than than 1.1 fm. From this one deduces that substantial contributions to the scattering come from ``short distance'' contributions which come from separations of greater than 1.1 fm. Recalling that $1/m_\pi \approx$ 1.4 fm, we see immediately that there is no significant scale separation between $m_\pi$ and the scales fixing the overall range of the nonpionic part of the potential. While the preceeding arguments do not decisively prove that $\Lambda \sim m_\pi$ they certainly show that it is not implausible. \section{$Q$ Counting and Cutoffs} Now the problem comes down to computing $p \cot (\delta )$. This can be done in PDS as in ref.~1. For the present purpose it is instructive to consider the cutoff calculation and we take our discussion from ref.~5. The essential physical idea in this approach is to implement the separation of long distance physics from short distance physics directly in configuration space. A radius, $R$, is introduced as a matching point between long and short distance effects; renormalization group invariance requires that physical quantities must be independent of $R$. It is important, however, that $R$ be chosen large enough so that essentially all of the effects of the short distance physics is contained within $R$. The potential is divided into the sum of two pieces, a short distance part which vanishes for $r>R$ and a long distance part which vanishes for $r<R$. At $R$, the information about short distance effects is entirely contained in the energy dependence of the logarithmic derivative (with respect to position) of the wave function at $R$. Thus, provided we can parameterize this information systematically, we can formulate the problem in a way which is insensitive to the details of the short distance part of the potential. This insensitivity to the details of the short distance physics is at the core of why effective field theory works. For $r>R$, the Schr\"odinger equation is solved subject to the boundary conditions at $R$. For s wave scattering, the wave function at $R$ may be parameterized as $A \sin (k r + \delta_0)$; the energy dependence of the logarithmic derivative is independent of $A$ and can be expressed in terms of an expansion similar to an effective range expansion: \begin{equation} p \cot (\delta_0) \, = \, -1/a_{\rm short} \, + \, 1/2 \,r_e^0\,p^2 \, + \, v_2^0 \, p^4 + \, v_3^0 \, p^6 + \,v_4^0 \, p^8 + \ldots \label{effrange0} \end{equation} Power counting in $Q$ for s wave scattering can be implemented straightforwardly. All of the coefficients in the preceding expansion are assumed to be order $Q^0$ except the first term ($-1/a_{\rm short}$ ) which will be taken to be order $Q^1$ to reflect the unnaturally large scale of the scattering length. Power counting for the long range part of the potential simply follows Weinberg's analysis \cite{Weinberg}, with the proviso that the potentials are only used for $r >R$. At order $Q^2$ in $p \cot (\delta )$, only the simple one-pion-exchange contribution to the $V_{\rm long}$ contributes. The power counting also justifies an iterative solution of the Schr\"odinger equation for $r>R$ along the lines of a conventional Born series. It differs from the usual Born series in that the boundary conditions at $R$ are implemented. Finally, $Q$ counting is used in expanding out the final expression for $k \cot (\delta )$. Carrying out this program gives the following expression for $k \cot (\delta )$ at order $Q^2$ for the ${}^1S_0$ channel \begin{eqnarray} p \cot (\delta ) & = &-\frac{1}{a_0} \, + \, m_\pi^2 \, \left[d + \, \frac{g_A^2 M}{16 \pi f_\pi^2} \, \left (\gamma + \ln (m_\pi R) \right) \right ] \, \nonumber \\ \nonumber \\ & + &\, \frac{1}{2} \, r_e^0 \,p^2 - \, \, \frac{g_A^2 M}{64 \pi a_0^2 f_\pi^2} \, \left( \frac{m_\pi^2}{p^2} \right ) \ln \left (1 + \frac{4 p^2}{m_\pi^2} \right ) \nonumber \\ \nonumber \\ & + & \, \, \frac{g_A^2 m_\pi M}{16 \pi a_0 f_\pi^2} \, \left( \frac{m_\pi}{p} \right )\, \tan^{-1} \left ( \frac{2 p}{m_\pi} \right ) \, + \, \frac{g_A^2 m_\pi^2 M}{64 \pi f_\pi^2} \, \ln \left (1 + \frac{4 p^2}{m_\pi^2} \right ) \label{kcotd1}\end{eqnarray} The convention used here has $f_\pi$ = 93 MeV. Apart from well-known parameters from pionic physics, there are three parameters---$a_0$, $d$ and $r_e^0$, where $1/a_{\rm short}$ from eq.~(\ref{effrange0}) is rewritten as $1/a_0 + d m_{\pi}^2$ with $1/a_0 \sim Q$, and $d m_\pi^2 \sim Q^2$. These parameters fix the energy dependence of the wave function at the matching scale $R$; renormalization group invariance requires $d$ to depend on R logarithmically. This form is precisely equivalent to the calculation in PDS, provided that the following identifications are made between the coefficients used above and those used in PDS with the notation of ref.~1. $$\frac{4 \pi}{M} \, \frac{1}{-\mu + 1/a_0} \, = \, C_0 $$ $$\frac{1}{2} \, r_e^0 \, = \, \frac{C_2 M}{4 \pi } \, \left ( \mu^2 \, - \, \frac{2 \mu}{a_0} \, + \, \frac{1}{a_0^2} \right ) $$ \begin{eqnarray} & m_\pi^2 &\, \left[d \, + \, \frac{g_A^2 M}{16 \pi f_\pi^2} \, \left (\gamma + \ln (m_\pi R) \right) \right] \, = \nonumber \\ \nonumber \\ \, & \frac{g_A^2 \, M}{16 \pi \, f_\pi^2} &\, \left ( m_\pi^2 \, \ln \left (\frac{m_\pi}{\mu} \right ) \, - \, m_\pi^2 \, + \, \frac{1}{a_0^2} \, - \, 2 \frac{\mu}{a_0} \, + \mu^2 \right ) \nonumber \\ \nonumber\\ & + &\, \frac{D_2 M}{4 \pi} \left ( m_\pi^2 \mu^2 \, - \, \frac{2 m_\pi^2 \mu}{a_0} \, + \, \frac{m_\pi^2}{a_0^2} \right) \label{equiv}\end{eqnarray} Formally, this is encouraging in the sense that it explicitly demonstrates the scheme independence of physical quantities. At the same time, there is an important hint of trouble which may lie ahead. In doing the derivation the matching scale, $R$ was taken to scale as $Q^0$ and the quantity $m_\pi R$ as order $Q^1$. To obtain the final expression only the leading term in $m_\pi R$ is kept. If $R \sim 1/m_\pi$ this is clearly problematic, and from the previous discussion about potentials, we see that $m_\pi R \sim 1$. It should also be stressed that the quantity $p \cot (\delta )$ is an extremely useful observable to work with in $Q$ counting. Unlike the amplitude itself, there are no poles near $p=0$; thus the issues of reorganizing the expansion as in OS \cite{MS} do not come up. Moreover, the expression is valid near $p=0$ (assuming that $Q$ counting holds) so it should be useful for ultra low energy scattering. \section{Low Energy Theorems} One difficulty with eq.~(\ref{kcotd1}) is that it is given in terms of $a_0$ which is the scattering length for the short distance potential only; as such it is not an observable. However, one can express everything in terms of physical observables and in doing so develop ``low energy theorems'' \cite{CH2}. The trick is to rel ate $a_0$ to the physical scattering length as follows: \begin{eqnarray} - \frac{1}{a} \, & = &\, - \frac{1}{a_0} + \, m_\pi^2 \, \left[ d \, + \, \frac{g_A^2 M}{16 \pi f_\pi^2} \, \left ( \gamma + \ln (m_\pi R) \right) \right ] \, + \, \frac{g_A^2 M}{16 \pi f_\pi^2} \, \left( \frac{2 m_\pi}{a_0} \, - \, \frac{1}{a_0^2} \right ) \nonumber \\ \nonumber \\ \, & = & \, - \,\frac{1}{a_0} + {\cal O}(Q^2/\Lambda) \label{scat} \end{eqnarray} Therefore in all of the ${\cal O}(Q^2)$ terms in eq.~(\ref{kcotd1}) one can replace $a_0$ by the physical $a$; the error in doing this is ${\cal O}(Q^3)$ which is one order beyond the order at which I am working. One gets the following expression orginally derived in ref.~8. \begin{eqnarray} p \cot (\delta ) & = &-\frac{1}{a_0} \, + \, m_\pi^2 \, \left[d + \, \frac{g_A^2 M}{16 \pi f_\pi^2} \, \left (\gamma + \ln (m_\pi R) \right) \right ] \, \nonumber \\ \nonumber \\ & + &\, \frac{1}{2} \, r_e^0 \,p^2 - \, \, \frac{g_A^2 M}{64 \pi a^2 f_\pi^2} \, \left( \frac{m_\pi^2}{p^2} \right ) \ln \left (1 + \frac{4 p^2}{m_\pi^2} \right ) \nonumber \\ \nonumber \\ & + & \, \, \frac{g_A^2 m_\pi M}{16 \pi a f_\pi^2} \, \left( \frac{m_\pi}{p} \right )\, \tan^{-1} \left ( \frac{2 p}{m_\pi} \right ) \, + \, \frac{g_A^2 m_\pi^2 M}{64 \pi f_\pi^2} \, \ln \left (1 + \frac{4 p^2}{m_\pi^2} \right ) \label{kcotd2}\end{eqnarray} One can expand this as a Taylor series in $p$ to obtain ERE coefficients. They are given by: \begin{eqnarray} v_2 \, & = & \, \frac{g_A^2 M}{16 \pi f_\pi^2} \, \left ( \, -\frac{16}{3 a^2 \, m_\pi^4}\, + \, \frac{32}{5 a \,m_\pi^3} \, - \,\frac{2}{m_\pi^2} \right )\nonumber \\ \nonumber \\ v_3 \, & = & \, \frac{g_A^2 M}{16 \pi f_\pi^2} \, \left ( \, \frac{16}{ a^2 \, m_\pi^6}\, - \, \frac{128}{7 a \, m_\pi^5} \, + \,\frac{16}{3 m_\pi^4} \right ) \nonumber \\ \nonumber \\ v_4 \, & = & \, \frac{g_A^2 M}{16 \pi f_\pi^2} \, \left ( \, -\frac{256}{5 a^2 \, m_\pi^8}\, + \, \frac{512}{9 a \,m_\pi^7} \, - \, \frac{16}{ m_\pi^6} \right ) \nonumber \\ \nonumber \\ & \ldots & \label{vi} \end{eqnarray} The preceding low energy theorems are valid to leading nontrivial order in $Q$ counting; corrections are of {\it relative} order $Q/\Lambda$. Several features of these expressions are notable. The first is that they are true predictions, independent of choices made in fitting, to this order in $Q$ counting. As such they are low energy theorems. One consequence of this is that different schemes ({\it eg.} fitting out the pole as in OS rather than using $1/a$) will only give corrections at relative order $Q/\Lambda$. In this sense these predictions can be considered ``low energy theorems '' which become exact in the limit $(m_\pi,1/a)/\Lambda \rightarrow 0$. The second significant fact is that all terms for all the expressions for the $v_i$ coefficients diverge in the chiral limit of $m_\pi \rightarrow 0$. This implies that these quantities are dominated by pionic effects and hence are a good place to test whether the pionic parts of $Q$ counting are working. If the pionic parts of $Q$ counting were under control one would expect that these predictions would work well. In practice, however, they work quite poorly. This can be seen in Table~(\ref{LET}) where the prediction from the low energy theorems are compared with coefficients extracted from the Nijmegen partial wave analysis (PWA). The prediction from the low energy theorems are typically off by a factor of 5 or so. This suggests that pionic parts of the $Q$ counting are failing rather badly. \begin{table}[t] \caption{A comparison of the predicted effective range expansion coefficients, $v_i$, for the ${}^1S_0$ and ${}^3S_1$ channels with coefficients extracted from the Nijmegen partial wave analysis.\label{LET}} \vspace{0.2cm} \begin{center} \footnotesize \begin{tabular}{|c| c | c | c |} \hline & $v_2$ (${\rm fm}^{3})$ & $v_3$ (${\rm fm}^{5})$ & $v_4$ (${\rm fm}^{7})$\\ \hline \hline $\delta$ (${}^1S_0$ channel)& & & \\ \hline \hline & & &\\ low energy theorem & -3.3 & 17.8 & -108. \\ \hline & & & \\ partial wave analysis & -.48 & 3.8 & -17. \\ \hline \hline $\delta$ (${}^3S_1$ channel)& & & \\ \hline \hline & & & \\ low energy theorem & -.95 & 4.6 & -25. \\ \hline & & & \\ partial wave analysis & .04 & .67 & -4.0\\ \hline \end{tabular} \end{center} \end{table} One possible difficulty with the comparison of the $v_i$ coefficients from the low energy theorems with the experimental data is that there is no experimental data. The coefficients extracted from the Nijmegen PWA were based on a fit to the smoothed ``best fit''. In principle one should do this fit including an error analysis based on the uncertainties. This cannot be done from the published data of the Nijmegen group as they did not publish information about correlated errors. Thus, one might wonder whether it is meaningful to extract high derivatives which are presumably rather sensitive to errors. A simple error estimate in ref.~4 concludes that the errors are likely to be too large to get any quantitative information about the $v$ coefficients. From this one might conclude the disagreement between the low energy theorems and the ``data'' in Table~(\ref{LET}) is due to an inability to extract the $v$ from the scattering data in a reliable way. This is almost certainly not the case, however. The Nijmegen group made several independent fits to the scattering data. One was the PWA. The others were various potential models which were fit directly to the data ({\it i.e.} not to the PWA). These fits had a $\chi^2$ per degree of freedom of essentially unity. Thus they can be regarded as independent fits to the data \cite{Nij2}. As the potential models have different forms from each other, they clearly have different systematic errors. Moreover in doing the least squares fit different models make different compromises in fitting individual data points so that they tend to explore the statistical errors. Thus, one might expect that the spread in the coefficients extracted in the different fits gives a reasonable feel for the scale of the uncertainty. Table~(\ref{pot}) shows the $v_i$ coefficients for these fits for the triplet channel \cite{Stoksa}, and it is manifestly clear that the spread in the effective range parameters as extracted from the three is vastly smaller than the difference with the predictions from the low energy theorems. \begin{table}[t] \caption{A comparison of the effective range expansion coefficients, $v_i$, for the ${}^3S_1$ and ${}^3S_1$ channel predicted from the low energy theorem with coefficients extracted from the partial wave analysis and with three potential models---Nijmegen I, Nijmegen II and Reid 93---which were fit directly to the scattering data. \label{pot}} \vspace{0.2cm} \begin{center} \begin{tabular}{|c| c | c | c |} \hline $\delta$ (${}^3S_1$ channel)&$v_2$ (${\rm fm}^{3})$ & $v_3$ (${\rm fm}^{5})$ & $v_4$ (${\rm fm}^{7})$\\ \hline \hline & & & \\ low energy theorem & -.95 & 4.6 & -25. \\ \hline & & & \\ partial wave analysis & .040 & .672 & -3.96\\ Nijm I & .046 & .675 & -3.97 \\ Nijm II & .045 & .673 & -3.95 \\ Reid93 & 0.033 & .671 & -3.90 \\ \hline \end{tabular} \end{center} \end{table} \section{ Re-summing the Effective Range Expansion} The effective range expansion parameters discussed in the previous section provide a dramatic way to see that the pionic parts of $Q$ counting may have serious problems with convergence. However, there are a number of drawbacks with looking at the $v_i$ coefficients. As noted above, there are ambiguities in the extraction from the data and it is hard to get reliable error bars. Moreover, the effective range expansion itself has a very limited radius of convergence. Because of the pion cut one expects the effective range expansion to converge only for $p < m_\pi /2$. Of course, all of the low energy theorems for the $v_i$ coefficients are contained in eq.~(\ref{kcotd2}). We can study this directly without expanding as a function of $k$. In effect, this amounts to re-summing the effective range expansion and using $p \cot (\delta )$ as our fundamental quantity. There are two obvious advantages to doing this: First, one can avoid the problem of extracting the $v_i$ coefficients from noisy data and instead we can compare directly with the partial wave analysis (which includes error estimates). Second, one is no longer restricted to $p < m_\pi /2$ since the re-summed expression is valid over the entire domain of $Q$ counting. Unfortunately, $p \,cot (\delta )$ does not isolate the pionic contributions from the rest, and at low $p$ it is dominated by $1/a$ physics and the fitting procedure which gives the effective range. There is a clean way to finesse this problem, however. Rather than study $p \cot (\delta)$ directly, one can study the following ``shape function''\cite{CH3}: \begin{equation} {\cal S}(p) \, = \, p \cot (\delta) - (-1/a + 1/2 r_e p^2) \end{equation} which is just the re-summed effective range expansion with the first two terms subtracted off. This has the advantage of removing completely the sensitivity to $1/a$ and the fitting of the effective range. The quantity is completely pion dominated since in a theory with pions integrated out ${\cal S} (p)$ is ${\cal O}(Q^3)$, while in a theory with explicit pions it is ${\cal O}(Q^2)$. In Table ~(\ref{S}) we see the low energy theorem prediction for ${\cal S} (p)$ compared with values extracted from the Nijmegen partial wave analysis for the triplet channel. Note the disagreement is quite pronounced. Moreover, note that error estimates are given for the results extracted from the scattering data. It is manifestly clear that the discrepancies are {\it not} due to uncertainties in the data. Again this suggests that the pionic parts of $Q$ counting are not predictive at leading nontrivial order---at least not for this observable. \begin{table}[t] \caption{A comparison of the shape function ${\cal S}(p^2) = p \cot (\delta) + 1/a - 1/2 r_e p^2 $ for the ${}^3 S_1$ channel extracted from the Nijmegen partial wave analysis with the prediction by the low energy theorem. \label{S}} \vspace{0.2cm} \begin{center} \begin{tabular}{|c |c | c|} \hline lab energy (MeV) &${\cal S}$ extracted (Mev) & ${\cal S}$ low energy theorem (Mev)\\ \hline \hline Deuteron Pole & $-0.0017 \pm 0.0125$ & -0.743\\ 1 & $-0.00095 \pm 0.00721$ & -0.0258 \\ 5 & $0.0428 \pm 0.0194$ & -0.535 \\ 10 & $0.245 \pm 0.047$ & -1.78 \\ 25 & $ 2.18 \pm 0.14$ & -7.54 \\ 50 & $ 11.03 \pm 0.24 $ & -20.10\\ \hline \end{tabular} \end{center} \end{table} \section{Conclusions} The development of $Q$ counting is an extremely important step in our understanding of effective field theory for nuclear physics. However, as stressed in this talk $Q$ counting involves two small mass parameters, $1/a$ and $m_\pi$. While there is every indication that the part of the theory based on expanding in $1/(a \Lambda)$ is working well, the chiral counting is far more problematic. The expansion has no predictive power for the effective range parameters and the shape function at leading nontrivial order (NLO). As both of these quantities are chirally sensitive this failure suggests that the chiral expansion may not be well under control. There are a number of possibilities. The most optimistic one is simply that the NLO calculation is not adequate, and if one works at higher order all will be well. The most pessimistic possibility is that that the chiral expansion is not converging and that this failure is general. Clearly, the way to resolve the situation is to work at higher order. It is important when doing so, however, to focus on observables which are highly sensitive to the pion physics. \section*{Acknowledgments} Most of the work discussed in this talk was done in collaboration with James Hansen. The research was funded by the U.S. Department of Energy under grant no. DE-FG02-93ER-40762. \section*{References}
\section*{ABSTRACT} The broad band 0.1-200 keV spectra of a sample of 5 Seyfert 2 galaxies (NGC 7172, Mkn 3, NGC 2110, NGC 4507 and NGC 7674) have been measured within the first year of the $BeppoSAX$ Core program. All sources have been detected up to $\sim$ 100 keV and their spectral characteristics derived with good accuracy. Although the results obtained from the detailed analysis of individual sources indicate some ``source-by-source'' differences, we show in the following that all spectra are consistent, at least qualitatively, with what expected from a ``0$^{th}$-order'' version of unified models. Indeed, a simple test on these data indicates that these Seyfert 2 galaxies are on average intrinsically very similar to Seyfert 1 galaxies (i.e., steep at E$\hbox{\raise0.5ex\hbox{$>\lower1.06ex\hbox{$\kern-1.07em{\sim}$}$}}$ 10 keV) and that the main difference can be ascribed to a different amount of absorbing matter along the line of sight (i.e. different inclinations of our line of sight with respect to a putative molecular torus or different thicknesses of the tori). \section{\bf PREDICTIONS FROM UNIFIED MODELS} The discovery of broad emission lines in the polarized optical spectra of several Seyfert 2 galaxies (Antonucci \& Miller 1985, Tran et al. 1992) has provided the basis for a unified model of Seyfert galaxies in which the main discriminating parameter between Seyfert 1 and Seyfert 2 nuclei is the inclination with respect to our line of sight of a supposed obscuring torus surrounding the central source (see Antonucci 1993 for a review). In this scheme, Seyfert 1 galaxies are active nuclei observed nearly perpendicularly to the torus plane (unabsorbed) whereas Seyfert 2 galaxies represent those seen through the torus (absorbed) and, of course, the main prediction is that Seyfert 2 galaxies are {\it intrinsically} similar to Seyfert 1 galaxies, once the effects of the torus are properly accounted for. It is now widely recognized that the {\it intrinsic} high energy spectrum of Seyfert 1 galaxies consists, on average, of a steep power-law continuum ($\Gamma$ $\sim$ 1.9--2.0, Nandra \& Pounds 1994) with an exponential cut-off typically at energies larger than 150 keV (Zdziarski et al. 1996, Perola et al. 1998). Therefore, one would expect that Seyfert 2 galaxies exhibit similar high-energy properties. X-ray observations (mainly below $\sim$ 20 keV) of Seyfert 2 galaxies, have shown a variety of spectral characteristics not always consistent with a ``0$^{th}$-order'' version of unified models (Smith \& Done 1996, Cappi et al. 1996, Turner et al. 1998). However, at high energies where the effects of absorption and matter reprocessing are less evident, measurements are sparse for Seyfert 2 galaxies (but see Johnson et al. 1997). It has been therefore difficult to assess the Seyfert 1 nature of the primary spectrum of Seyfert 2 galaxies from these observations. Based on these general considerations, we have undertaken a program of observations with $BeppoSAX$, aimed at studying the X-ray spectral properties of bright Seyfert 2 galaxies over a broad energy band (up to about 200 keV) and at testing the validity of unified models. As a matter of fact, $BeppoSAX$ can provide crucial information on the intrinsic source properties because high energy photons from a few to several keV can pass through the circumnuclear intervening material and can therefore be compared to the ``typical'' spectrum of Seyfert 1 galaxies. So far 5 objects have been observed within the first AO cycle, namely NGC 7674 (Malaguti et al. 1998a), Mkn 3 (Cappi et al.1998), NGC 2110 (Malaguti et al. 1998b), NGC 7172 and NGC 4507 (see also Bassani et al. 1998). \section{A SIMPLIFIED QUALITATIVE TEST} A simple, qualitative, test has been performed on our sources by fitting each source of the sample with the same model: a soft power-law component plus a hard, absorbed, power-law component with reflection and associated Fe K line as illustrated in Fig. 1. In the framework of unified models, the soft power-law is attributed to scattered soft X-ray emission from ionized material placed above the molecular torus, while the hard X-ray and reflection components are interpreted as the direct component absorbed by the torus and its reflection from the inner side of the torus, respectively. The intensity of the soft, scattered, component was assumed to be $\sim$ 2\% that of the direct one. The intensity of the reflection component was fixed to R = 1 (conrresponding to a 2$\pi$ coverage as viewed from the X-ray source), and the iron line was assumed to be produced through both the reflection and absorption with an equivalent width of $\sim$ 1 keV (with respect to the reflected continuum) and $\sim$ 500 $\times$ $N_{\rm H}\over{1.23 \times 10^{24}}$ eV (with respect to the direct, absorbed, component), as expected from theoretical models (George \& Fabian 1991, Leahy \& Creighton 1993). The only free parameters were the intensity and photon index of the direct continuum and the absorption column along the line of sight. The fit results and unfolded spectra obtained from this test are given in Figure 2. The most interesting result is that all sources are well described by a steep, Seyfert 1-like spectrum, with $\Gamma$ $\sim$ 1.79--1.95. This is a newly discovered behaviour of Seyfert 2 galaxies for energies up to $\sim$ 100 keV which supports unified models. It is stressed that this result is largely independent on the presence of the steep soft component and on the assumed intensity of the reflection component. This is demonstrated by the fact that the average Seyfert 2 spectrum obtained averaging all the PDS 20-200 keV data (except NGC 7674) can be well fitted ($\chi^2$ $\sim$ 1.2) by a single power-law model with $\Gamma_{20--200 keV}$ = 1.85 $\pm$ 0.05 and shows no deviation from the power-law up to $\sim$ 150 keV. The observed major differences in the quality of fits ($\chi^{2}_{red}$ ranging from 0.9--1.8) are to be ascribed to two main factors. The first is source by source differences in the Fe K complex which indicates the need of a more dedicated analysis. The second depends upon the different amount of absorbing matter along the line of sight. The less absorbed source is NGC 2110 which also shows less indication for a reflection component and the most absorbed one is NGC 7674, where only the reflection component is observed possibly because the direct component is completely blocked by a Compton thick molecular torus (with $N_{\rm H} \hbox{\raise0.5ex\hbox{$>\lower1.06ex\hbox{$\kern-1.07em{\sim}$}$}} 10^{25}$ cm$^{-2}$, Malaguti et al. 1998a). Intermediate cases are NGC 7172, NGC 4507 and Mkn 3; in the latter, both the direct and reflected components are clearly resolved spectroscopically. Moreover although a detailed measurement in individual sources of the high-energy cutoff is difficult, it appears clear in the data (see Figure 2) that there is no evidence of such a cutoff for energies up to at least $\sim$ 100--150 keV. {\bf REFERENCES} \vspace{-5mm} \begin{itemize} \setlength{\itemindent}{-8mm} \setlength{\itemsep}{-1mm} \item [] Antonucci, R.R.J., 1993, $ARA\&A$, {\bf 31}, 473 \item [] Antonucci, R.R.J. \& Miller, J.S., 1985, {\it ApJ}, {\bf 297}, 621 \item [] Bassani, L., et al. 1998, to appear in proceedings of ``Dal nano- al tera -eV: tutti i colori degli AGN'', third Italian conference on AGNs, Roma, {\it Memorie S.A.It}, astro-ph/9809327 \item [] Cappi M., Mihara, T., Matsuoka, et al., 1996, {\it ApJ}, {\bf 456}, 141 \item [] Cappi, M., et al. 1998, $A\&A$, in press, astro-ph/9902022 \item [] George, I.M. \& Fabian, A.C., 1991, {\it MNRAS}, {\bf 249}, 352 \item [] Johnson, W.N., Zdziarski, A.A., Madejski, G.M., Paciesas, W.S., Steinle, H., \& Lin Y-C, 1997, in proceedings of the Fourth Compton Symposium, Ed. D. Dermere, M.S. Stcikman and J.D. Kurfess, {\it AIP}, 283 \item [] Leahy D.A. and Creighton J., 1993, {\it MNRAS}, {\bf 263}, 314 \item [] Malaguti, G., et al. 1998a, $A\&A$, {\bf 331}, 519 \item [] Malaguti G. et al. 1998b, $A\&A$, in press, astro-ph/9901141 \item [] Nandra, K. \& Pounds, K.A., 1994, {\it MNRAS}, {\bf 268}, 405 \item [] Perola, C., et al. 1998, to appear in proceedings of ``Dal nano- al tera -eV: tutti i colori degli AGN'', third Italian conference on AGNs, Roma, {\it Memorie S.A.It} \item [] Smith, D.A., \& Done, C., 1996, {\it MNRAS}, {\bf 280}, 355 \item [] Tran, H.D., Miller, J.S. \& Kay, L.E., 1992, {\it ApJ}, {\bf 397}, 452 \item [] Turner, T.J., George, I.M., Nandra, K., \& Mushotzky, R.F., 1998, {\it ApJ}, {\bf 493}, 91 \item [] Zdziarski, A.A., Johnson, W.N., Poutanen, J., Magdziarz, P., \& Gierlinski, M., 1996, in ``The Transparent Universe'', proceedings of the 2nd INTEGRAL Workshop, Ed. C. Winkler, T.J.-L. Courvoiser and P. Durouchoux, ESA SP-382, 373 \end{itemize} \begin{figure}[htb] \psfig{file=./cappi_fig1.ps,width=16cm,height=8cm,angle=-90} \caption{Simple, qualitative, test model used to fit all the sources of the sample in the framework of unified models.} \end{figure} \par\noindent \vfill\eject \normalsize \begin{figure}[htb] \hspace{1.5truecm} \parbox{5truecm} {{\bf NGC 2110}: \par\noindent $N_{\rm H}$ $\sim$ 5 $\times$ 10$^{22}$ cm$^{-2}$ \par\noindent $\Gamma$ $\sim$ 1.97 \hspace{2.5truecm}{$\Longleftrightarrow$} \par\noindent $\chi^{2}_{red}$ $\sim$ 1.16 \par\noindent Malaguti et al. 1998, submitted to A\&A} \ \hspace{0truecm} \ \parbox{8truecm} {\psfig{file=./cappi_fig2.ps,width=8cm,height=4cm,angle=-90}} \par\noindent \hspace{1.5truecm} \parbox{5truecm} {{\bf NGC 7172}: \par\noindent $N_{\rm H}$ $\sim$ 10$^{23}$ cm$^{-2}$ \par\noindent $\Gamma$ $\sim$ 1.83 \hspace{2.5truecm}{$\Longleftrightarrow$} \par\noindent $\chi^{2}_{red}$ $\sim$ 1.4 \par\noindent Dadina et al., in prep} \ \hspace{0truecm} \ \parbox{8truecm} {\psfig{file=./cappi_fig3.ps,width=8cm,height=4cm,angle=-90}} \par\noindent \hspace{1.5truecm} \parbox{5truecm} {{\bf NGC4507}: \par\noindent $N_{\rm H}$ $\sim$ 5 $\times$ 10$^{23}$ cm$^{-2}$ \par\noindent $\Gamma$ $\sim$ 1.83 \hspace{2.5truecm}{$\Longleftrightarrow$} \par\noindent $\chi^{2}_{red}$ $\sim$ 1.8 \par\noindent Bassani et al., in prep} \ \hspace{0truecm} \ \parbox{8truecm} {\psfig{file=./cappi_fig4.ps,width=8cm,height=4cm,angle=-90}} \par\noindent \hspace{1.5truecm} \parbox{5truecm} {{\bf Mkn 3}: \par\noindent $N_{\rm H}$ $\sim$ 1.3 $\times$ 10$^{24}$ cm$^{-2}$ \par\noindent $\Gamma$ $\sim$ 1.79 \hspace{2.5truecm}{$\Longleftrightarrow$} \par\noindent $\chi^{2}_{red}$ $\sim$ 0.9 \par\noindent Cappi et al., submitted to A\&A} \ \hspace{0truecm} \ \parbox{8truecm} {\psfig{file=./cappi_fig5.ps,width=8cm,height=4cm,angle=-90}} \par\noindent \hspace{1.5truecm} \parbox{5truecm} {{\bf NGC 7674}: \par\noindent $N_{\rm H}$ $\hbox{\raise0.5ex\hbox{$>\lower1.06ex\hbox{$\kern-1.07em{\sim}$}$}}$ 10$^{25}$ cm$^{-2}$ \par\noindent $\Gamma$ $\sim$ 1.95 \hspace{2.5truecm}{$\Longleftrightarrow$} \par\noindent $\chi^{2}_{red}$ $\sim$ 0.9 \par\noindent Malaguti et al. 1998a} \ \hspace{0truecm} \ \parbox{8truecm} {\psfig{file=./cappi_fig6.ps,width=8cm,height=4cm,angle=-90}} \caption{Fit results and broad-band unfolded spectra obtained from the qualitative test. See text for a description of the model fitted. $N_{\rm H}$ increases going from the top (NGC 2110) to the bottom (NGC 7674).} \end{figure} \end{document}
\section{Introduction} Optimization methods have found widespread application in computational physics. Among these the investigation of the low-temperature behavior of spin glasses \cite{binder86} attracted most of the attention within the statistical physics community. The reason is that despite its simple definition (see below) its behavior is far from being understood. From the computational point of view the calculation of spin-glass ground states is very demanding, because it belongs to the class of the NP-hard problems \cite{barahona82}. This means that only algorithms are available, for which the running time on a computer increases exponentially with the system size. In this work a method recently proposed, the {\em cluster-exact approximation} (CEA) \cite{alex2} is applied to four-dimensional Ising spin glasses. The model under investigation here consists of $N$ spins $\sigma_i = \pm 1$, described by the Hamiltonian \begin{equation} H \equiv - \sum_{\langle i,j\rangle} J_{ij} \sigma_i \sigma_j \end{equation} where $\langle \ldots \rangle $ denotes a sum over pair of nearest neighbors. In this report simple 4d lattices are considered, i.e. $N=L^4$. The nearest neighbor interactions (bonds) take independently $J_{ij} = \pm 1$ with equal probability. Periodic boundary conditions are applied to the systems. No kind of external magnetic field is present here. Four-dimensional Ising spin glasses have been investigated rather rarely. Most of the results were obtained via Monte-Carlo (MC) simulations at finite temperature, see e.g. \cite{bhatt85u88,reger90,badoni93,parisi96,bernardi97,marinari98,hukushima99}. Here the $T=0$ behavior is investigated, i.e. ground states are calculated. This has the advantage, that one does not encounter ergodicity problems or critical slowing down like in algorithms which base on MC methods. Only one attempt \cite{wanschura96} to address the 4d spin-glass ground-state problem is known to the author. But, as we will see later, the former results suffer from the problem, that not the true global minima of the energy were obtained. Furthermore, no analytic predictions of the ground-state energy have been noted by the author. The question whether finite-dimensional Ising spin glasses show an ordered phase below a non-zero transition temperature $T_c$ is of crucial interest. By MC simulations around the (expected) transition temperature this question is hard to solve. Another way to address this question is to calculate the {\em stiffness} or {\em domain wall energy} $\Delta=E^a-E^p$ which is the difference between the ground-state energies $E^a, E^p$ for antiperiodic and periodic boundary conditions in one direction\cite{bray84,mcmillan84}. Here the antiperiodic boundary conditions for calculating $E^a$ are realized by inverting one plane of bonds. For the other directions periodic boundary conditions are applied always. This treatment introduces a domain wall into the system. If a model exhibits an ordered low-temperature phase, the domain wall increases with growing system size, which becomes visible through the behavior of $\Delta$: the disorder-averaged stiffness energy shows a finite size dependence \begin{equation} \langle |\Delta| \rangle \sim L^{\Theta_S} \end{equation} A positive value of the stiffness exponent $\Theta_S$ indicates the existence of an ordered phase for non-zero temperature. For example a simple $d=2$ Ising ferromagnet has $\Theta_S=1$. For spin glasses, the stiffness exponent plays additionally an important role within the droplet-scaling theory \cite{mcmillan,bray,fisher86,fisher88,bovier}, where it describes the finite-size behavior of the basic excitations (the droplets). Using this kind of analysis is was proven that the 2d spin glass exhibits no ordering for $T>0$ \cite{kawashima97}. For the three-dimensional problem in a recent calculation \cite{alex-stiff} by applying genetic CEA a value of $\Theta_S=0.19(2)$ was found, which shows, that indeed the $d=3$ model has a spin-glass phase for nonzero temperature. For $d=4$ the existence of a finite $T_c\approx 2.1$ was proven rather early even by MC simulations \cite{bhatt85u88,reger90}, but the value for the stiffness-exponent $\Theta_S$ is of interest on its own. In \cite{hukushima99} recently a value of $\Theta_S=0.82(6)$ was found by performing a MC simulation near $T_c$. In the work presented here the value is obtained via ground-state calculations. The paper is organized as follows: In the next section the algorithm applied here is briefly presented. The main section contains the results for the ground-state energy and the stiffness exponent. Finally a summary is given. \section{Algorithm} The technique for the calculation bases on a special genetic algorithm \cite{pal96,michal92} and on cluster-exact approximation \cite{alex2} which is an optimization method designed especially for spin glasses. Now a brief description of the method is given. Genetic algorithms are biologically motivated. An optimal solution is found by treating many instances of the problem in parallel, keeping only better instances and replacing bad ones by new ones (survival of the fittest). The genetic algorithm starts with an initial population of $M_i$ randomly initialized spin configurations (= {\em individuals}), which are linearly arranged in a ring. Then $\nu \times M_i$ times two neighbors from the population are taken (called {\em parents}) and two offspring are created using the so called triadic crossover \cite{pal95}. Then a mutation with a rate of $p_m$ is applied to each offspring, i.e. a fraction $p_m$ of the spins is reversed. Next for both offspring the energy is reduced by applying CEA. The algorithm bases on the concept of {\em frustration} \cite{toulouse77}. The method constructs iteratively and randomly a non-frustrated cluster of spins, whereas spins with many unsatisfied bonds are more likely to be added to the cluster. The non-cluster spins act like local magnetic fields on the cluster spins. For the spins of the cluster an energetic minimum state can be calculated in polynomial time by using graph-theoretical methods \cite{claibo,knoedel,swamy}: an equivalent network is constructed \cite{picard1}, the maximum flow is calculated \cite{traeff,tarjan} and the spins of the cluster are set to the orientations leading to a minimum in energy. This minimization step is performed $n_{\min}$ times for each offspring. Afterwards each offspring is compared with one of its parents. The pairs are chosen in the way that the sum of the phenotypic differences between them is minimal. The phenotypic difference is defined here as the number of spin where the two configurations differ. Each parent is replaced if its energy is not lower (i.e. better) than the corresponding offspring. After this creation of offspring is performed $\nu \times M_i$ times the population is halved: From each pair of neighbors the configuration which has the higher energy is eliminated. If not more than 4 individuals remain the process is stopped and the best individual is taken as result of the calculation. The whole algorithm is performed $n_R$ times and all configurations which exhibit the lowest energy are stored, resulting in $n_g$ statistical independent ground state configurations. The method was already applied for the investigation of the ground-state landscape of 3d Ising spin glasses \cite{alex-3d}. \section{Results} In this section at first the values for the simulation parameters, which are defined above, are presented. Then the finite-size behavior of the ground-state energy is investigated. Finally results for the stiffness energy are discussed. The simulation parameters were determined in the following way: For the system sizes $L=2,4,6,7$ several different combinations of the parameters $M_i, \nu, n_{min}, p_m$ were tested. For the final parameter sets it is not possible to obtain lower energies even by using parameters where the calculation consumes four times the computational effort. For $L=3,5$ the parameter sets for $L+1$ were used. Using parameter sets chosen this way genetic CEA calculates true ground states, as shown in \cite{alex-stiff}. It should be pointed out that it is relatively easy to obtain states, which exhibit an energy slightly above the true ground state energy. The hard task is to obtain really the global minimum of the energy. Here $p_m=0.1$ and $n_R=5$ were used for all system sizes. Table \ref{tab_parameters} summarizes the parameters. Also the typical computer time $\tau$ per ground state computation on a 80 MHz PPC601 is given. Ground states were calculated for system sizes up to $7\times 7\times 7 \times 7$ for $N_L$ independent realizations (see table \ref{tab_parameters}) of the random variables. For each realization the ground states with periodic and antiperiodic boundary condition in one direction were calculated. The remaining three directions are always subjected to periodic boundary conditions. One can extract from the table that for small system sizes $L\le 4$ ground states are rather easily to obtain, while the $L=7$ systems alone required 6560 CPU-days. Using these parameters on average $n_g>2.7$ ground states were obtained for every system size $L$ using $n_R=5$ runs per realization. The average ground-state energy $e_0$ per spin is shown in Fig.\ref{figEnergy} as a function of the system size $L$. Using a fit to $e_0(L)=e_0^{\infty} + a*L^{-b}$ the value for the infinite system is extrapolated, resulting in $e_0^{\infty} = -2.095(1)$ ($a=7.1(7),b=-4.2(1)$). This value is compatible with the lower bound of $e_0=\sqrt{2d\ln 2}\approx 2.35$ given by the random energy model \cite{derrida81}. The value calculated here is substantially smaller than the result $e_0^{\infty} = -2.054(3)$, which was obtained in \cite{wanschura96} using a pure genetic algorithm. This shows that in \cite{wanschura96} not the true global minima were found, which can be concluded also from the fact, that there $e_0(L)$ increases with growing system size. Because the periodic boundary conditions impose additional constraints on the systems, the opposite behavior is expected, as found for the results presented here. For further comparison additionally some calculations were performed by the author by simply rapidly quenching from random chosen spin configurations. By executing an analogous fit, a value of $e_0^{\infty} = -2.04(2)$ is obtained. This shows, that the result from \cite{wanschura96} seems to be only slightly better than the data obtained by applying a very simple minimization method. The distribution of the stiffness energy, which is obtained from performing ground-state calculations for systems with either periodic or antiperiodic boundary conditions in one direction, are shown in Fig. \ref{figPStiff} for $L=5$ and $L=7$. With increasing system size the distribution broadens. This means that larger domain walls become more and more likely. To study this effect more quantitatively, in Fig. \ref{figStiffness} the disorder-averaged absolute value $\langle |\Delta| \rangle$ of the stiffness energy is plotted as a function of the system size $L$. Also shown is a fit $\langle|\Delta(L)|\rangle \sim L^{\Theta_S}$ which results in $\Theta_S=0.64(5)$. Here, the system sizes $L=2,3$ were left out of the analysis, since they are below the scaling regime. Because of the large sample sizes the error bars are small enough, so we can be pretty sure that $\Theta_S>0$. It confirms earlier results from MC simulations \cite{bhatt85u88,reger90} that the 4d EA spin glass exhibits a non-zero transition temperature $T_c$. The value $\Theta_S=0.64(5)$ is comparable to a recent result from MC simulations $\Theta_S=0.82(6)$ \cite{hukushima99}, given the facts that the system sizes are rather small and the other result was obtained at finite temperature near the transition point $T_c\approx 2.1$. Additionally, the prediction from droplet-scaling theory $\Theta_S < (d-1)/2 = 1.5$ \cite{fisher88} is fulfilled. It should be pointed out, that the method described above does not guaranty to find exact ground states, although the method for choosing the parameters makes it very likely. If states with a slightly higher energy are obtained, the result for $e_0^{\infty}$ is not affected very much. For the stiffness energy, it was shown in \cite{alex-stiff} that the result is very reliable as well, as long as the energies of the states are not too far away from the true ground-state energies. \section{Conclusion} Results have been presented from calculations of a large number of ground states of 4d Ising spin glasses. They were obtained using a combination of cluster-exact approximation and a genetic algorithm. Using a huge computational effort it was ensured that true ground states have been obtained with a high probability. The finite size behavior of the ground-state energy and the stiffness energy have been investigated. By performing a $L\to\infty$ extrapolation, the ground-state energy per spin for the infinite system is estimated to be $e_0^{\infty}=-2.095(1)$. The absolute value of the stiffness energy increases with system size and shows a $\langle |\Delta(L)|\rangle\sim L^{\Theta_S}$ behavior with $\Theta_S=0.64(5)$. For systems with a Gaussian distribution of the bonds qualitatively similar results are expected, since the ordering behavior depends only on the sign of the interactions and not on their magnitudes. A more detailed study of the ground-state landscape of 4d systems, similar to \cite{alex-3d}, requires more than $n_G\approx 3$ ground states per realization to be calculated. Since this requires a substantial higher computational effort, it remains to be done for the future. \section{Acknowledgements} The author thanks A.P. Young for interesting discussions, critical reading of the manuscript and for the allocation of computer time on his workstation cluster at the University of California in Santa Cruz. This work was suggested by him during the ``Monbusho Meeting'' held at the {\em Fondation Royaumont} near Paris. The author was supported by the Graduiertenkolleg ``Modellierung und Wissenschaftliches Rechnen in Mathematik und Naturwissenschaften'' at the {\em In\-ter\-diszi\-pli\-n\"a\-res Zentrum f\"ur Wissenschaftliches Rechnen} in Heidelberg and the {\em Paderborn Center for Parallel Computing} by the allocation of computer time. Financial support was provided by the DFG ({\em Deutsche Forschungsgemeinschaft}) and the organizers of the ``Monbusho Meeting''.
\section*{Acknowledgments} We are grateful to Manuel Drees and Peter Zerwas for valuable comments and helpful discussions. This work was supported in part by the Korea Science and Engineering Foundation (KOSEF) through the KOSEF--DFG large collaboration project, Project No.~96--0702--01-01-2, and in part by the Center for Theoretical Physics. MG acknowledges Alexander von Humboldt Stiftung foundation for financial help and also KOSEF for funding during his stay in Yonsei University, Seoul, where this work was initiated.
\section{Introduction} In order to test models for structure formation in the Universe it is necessary to pinpoint galaxies within the modelled large-scale distribution of matter. As this distribution evolves in a non-linear fashion, the use of N-body methods is usually required. Galaxies should form, under the influence of gravity, within an N-body simulation in a self-consistent way. However, such simulations suffer from a numerical problem: small groups of particles that represent galaxies get disrupted by numerical two-body effects within clusters (Carlberg 1994; van Kampen 1995). This can be solved by replacing each group of particles by a single `galaxy particle' just after they have formed into a virialised system that resembles a galactic halo, thus ensuring their survival. This should also produce galaxies at the right time and the right place, with a spectrum of masses. In van Kampen (1995) such galaxy particles were only formed at a single epoch. Here we extend this scheme to `continuous' galaxy formation by applying the algorithm several times during the evolution. This means that merging of already-formed galaxies is taken into account as well, although only in a schematic fashion. An important advantage of having a galaxy formation algorithm added to the N-body integrator is that the time normalisation is now intrinsically fixed, since one can {\it directly} compare the properties of the distribution of simulated galaxies to those of the observed galaxy distribution. \section{Galaxy formation recipe} \subsection{Outline} For the identification of galaxies during the evolution of the large-scale matter distribution we use the {\it local density percolation} algorithm (see also van Kampen 1995). Instead of a fixed linking length, we modulate the linking length according to the {\it local density}\ around each particle in such a way that the linking length is shorter in high-density environments. This partly resolves the cloud-in-cloud problem. We form (and merge) galaxies several times during the evolution, so we have to consider percolation of {\it unequal}\ mass particles. In the following ${\bf x}$ and ${\bf v}$ are {\it comoving} variables. \subsection{Local density percolation for unequal mass particles} In the local density percolation scheme particles are linked together if they are separated by a certain fixed fraction $p$ of the Poissonian average nearest neighbour distance $x^{\rm P}_{\rm nn}\equiv [4\pi\langle n\rangle /3]^{-1/3}$, where $\langle n\rangle$ is the mean number density of particles, modulated by the local number density $n^{\rm G}({\bf x},s)$, which is $n({\bf x})$ Gaussian smoothed at the scale $s\ x^{\rm P}_{\rm nn}$. Thus, $p$ and $s$ are the (dimensionless) free parameters of the algorithm. All models in this paper have $\langle n \rangle=8.0\ h_0^3$Mpc$^{-3}$, so $x^{\rm P}_{\rm nn}=0.31 h_0^{-1}$Mpc. For unequal mass particles we need an extra modulation according to their masses. Initially, all particles have mass $m_0$. As galaxies form, particles arise with masses $m_i$ that are integer multiples of $m_0$. Galaxy particles with mass $m_i$ that are put $\sqrt{m_i/m_0}$ times further away will exert the same gravitational force as dark particles with mass $m_0$. So we should take the factor $\sqrt{m_i/m_0}$ as the second modulation factor. This gives a local percolation length $$ R_{\rm p}({\bf x}_i,p,s) = p\ x^{\rm P}_{\rm nn} \sqrt{m_i\over m_0} \Bigl[{n^{\rm G}({\bf x}_i,s)\over\langle n\rangle}\Bigr]^{-{1/3}}\ .\eqno(1)$$ Since each particle has its own linking length, we use their mean to test pairs of particles. Furthermore, to prevent excessive percolation lengths, we adopt an absolute maximum of $x^{\rm P}_{\rm nn}/2$ for the pair linking lengths (after taking the mean of the individual ones), i.e.\ a lower limit for the local particle density which is equal to eight times the background particle density. In addition, we require pairs to have a relative pairwise velocity $$ v_{||} \equiv({\bf v}_j - {\bf v}_i)\cdot({\bf x}_j - {\bf x}_i) / |{\bf x}_j - {\bf x}_i| \eqno(2) $$ of less then 800 km s$^{-1}$. This is twice the relative pairwise velocity dispersion at separations around $x^{\rm P}_{\rm nn}$ for the field, and somewhat smaller than found for a sample including the Coma cluster (Mo, Jing \& B\"orner 1993). It is also twice the maximum internal velocity dispersion we allow for a group. We include this velocity linking length to exclude fast-moving particles which are geometrically linked to a group. This often occurs within the potential wells of galaxy clusters. One should see this criterion as the velocity equivalent of a (constant) spatial linking length, so that we actually find groups in phase-space. However, the velocity linking length is less restrictive than the spatial linking length since it serves a different purpose, as said. \subsection{Virial equilibrium criterion for unequal mass particles} A group of particles should only be transformed into a single, soft galaxy particle if it forms a physical system roughly in virial equilibrium. This `virial criterion' is a necessary addition to the local density percolation algorithm for the purpose of defining galaxies. We will use a virial equilibrium criterion in a simplified form using the half-mass radius $R_{\rm h}$, motivated by Spitzer (1969) who found that for many equilibrium systems the virial equilibrium equation can be written as (where $v$ is now the proper velocity) $$\sigma_v\equiv\langle v^2 \rangle \approx 0.4 {GM\over R_{\rm h}}\ . \eqno(3)$$ If galaxies are identified only once, one has to deal with equal mass particles, and $R_{\rm h}$ can simply be calculated by obtaining the median of the distances of all particles with respect to the centre of the group being tested. For groups where the masses of the member particles can differ a few orders of magnitude, the half-mass radius will often exactly coincide with a galaxy particle. This makes the median (i.e.\ the half-mass radius) a rather noisy estimater for the total gravitational energy of a group of unequal mass particles. Because for many probability distributions the mean and the median are almost identical, we use the mass-weighted mean distance from the centre of the group as an estimator for $R_{\rm h}$. This is a more smoothly-defined and well-behaved quantity than the median distance. The new galaxy particle has a softening parameter corresponding to the $R_{\rm m}$ of the original group, which certifies a reasonable conservation of energy (see van Kampen 1995). Discreteness noise will cause some scatter in the group quantities, so we should allow for some tolerance in the difference between the estimated virial mass and the true mass of the group. The allowed tolerance determines the `reach' the criterion has in time: larger permitted deviations from virial equilibrium result in the acceptance of groups that are still collapsing. We accept groups as real when the virial mass is within 25 per cent of the true mass. \subsection{Choice of the galaxy formation parameters} For the (dimensionless) local density percolation parameters we choose $p=1$ and $s={\scriptstyle 1 \over 2}$. The maximum percolation length is $x^{\rm P}_{\rm nn}/2$, which is $0.16 h_0^{-1}$Mpc for our simulations. We set the upper mass limit for galaxy particles to be $1.4\times10^{13}$ M$_\odot$. This ensures that possible cD galaxies are {\it not}\ modelled by single galaxy particles, since that would produce undesirable numerical problems, and galaxies that massive are not (numerically) disrupted anyway. We add an extra limit on the internal galaxy velocity dispersion of $\sigma_v < 400$ km s$^{-1}$, the maximum value found for typical ellipticals (de Zeeuw \& Franx 1991) and comfortably within the velocity dispersion of haloes around spirals given their typical circular velocities of 200-300 km s$^{-1}$. Finally, we need to adopt a lower limit of seven particles in a group because of discreteness noise that causes an artificially large scatter in the virial mass estimate. \section{Description and timing of the simulations} We have run eight simulations of average patches of universe, and 99 cluster models. This latter set forms a catalogue of galaxy clusters, and is discussed extensively in van Kampen \& Katgert (1997). The actual N-body code we use is the Barnes \& Hut (1986) treecode, slightly adapted for our purposes and supplemented with the galaxy formation algorithm. We ran the models up to $\sigma_8=1$, which is sufficiently beyond the time that is expected to be the present epoch for the $\Omega_0=1$ CDM scenario adopted: $\sigma_8$ was found to be significantly smaller than unity in most earlier work (e.g.\ Davis et al.\ 1985; Frenk et al.\ 1990; Bertschinger \& Gelb 1991). From a comparison of the galaxy-galaxy autocorrelation function obtained for the field models to that observed, we find $\sigma_8$ to be in the range 0.46 to 0.56 (van Kampen 1997), while a similar comparison of the statistical properties of clusters gives roughly the same range (van Kampen \& Katgert 1997). \section{Galaxy properties} \subsection{Galaxy formation and merging rates} As a first check how our modelling of the formation and merging of galaxies compares to other techniques, notably hydrodynamical simulations, we look at the galaxy formation and merger number density rates as a function of time. These are plotted in Figure~\ref{fig-1} for $\sigma_8=0.46$, where $t_0$ is the present epoch. The formation rate peaks at $z\approx 1.3$, whereas the merger rate does not show a clear peak. The merging of small objects into galaxies with masses that are included in the formation rate is not included in the merger rate. \begin{figure} \psfig{file=frates.ps,width=10.0cm,silent=} \caption{Galaxy formation and merger rates for our models (symbols) and the simulations of Summers (1993, histograms), rescaled to our units and galaxy masses (see text). Filled symbols and solid lines represent galaxy formation, open symbols and dotted lines merging.} \label{fig-1} \end{figure} The shapes and amplitudes of both rates compare remarkably well with those found from hydrodynamical simulations performed by Summers (1993), also shown in Figure~\ref{fig-1}, if we triple his time-scale. This can be justified quantitatively as follows: Summers (ibid.) has a higher mass resolution and forms galaxies down to a lower mass cut-off. The smallest galaxy masses in his simulations are roughly a hundred times smaller than our lowest mass galaxies. The CDM spectrum on galactic scales is a power-law with index -2. We can then use the scaling law $t_{\rm form}\sim M^{1/4}$ which applies for such a spectrum to find that this mass difference gives a factor of three difference in the formation time. Summers (ibid.) used a full-fledged hydro code (and identified galaxies with the ordinary friends-of-friends algorithm). The fact that we find similar shapes and amplitudes for the rates means that the use of a galaxy formation recipe with an ordinary collisionless N-body code can give comparible results to more advanced simulation techniques that incorporate more (but certainly not all) physical processes. \subsection{Cluster luminosity function} For our cluster simulations, we study the luminosity function within the projected Abell radius. Since we know only the masses of the galaxies we need to assume a constant mass-to-light ratio $\Upsilon$ to obtain a luminosity function. For the $B_{\rm J}$ magnitude, $\Upsilon_{\rm J}\approx1200$ for an $\Omega_0=1$ universe. Because on average 25 per cent of the mass is locked into our galaxies (including dark haloes), $\Upsilon_{\rm J}\approx 300$ for the galaxies. The joint luminosity function for our cluster models is plotted in Figure~\ref{fig-2} for all galaxies (all symbols), and for the $M>1.5\times10^{12}$ M$_\odot$ ones (filled symbols only), along with a fitted Schechter (1976) luminosity function for each of them (dashed line for all galaxies, solid line for the limited set), with slope $\alpha$ and characteristic magnitude $B_{\rm J}^*$ as free parameters. \begin{figure} \psfig{file=mf1a.ps,width=10.0cm,silent=} \caption{Joint luminosity function for all galaxies within the projected Abell radius. Filled symbols represent galaxies with masses larger than $1.5\times10^{12}$M$_\odot$. The dashed line indicates a fit of a Schechter (1976) function to all galaxies. The solid line is a similar fit to the massive ones only (filled symbols). Both fits were made with $\alpha$ and $B_{\rm J}^*$ as free parameters, and assume a constant mass-to-light ratio $\Upsilon_{\rm J}=300$.} \label{fig-2} \end{figure} The fit for all galaxies is quite good but not perfect: $\alpha=-1.5$. We find that the fit gets better for the limited set of massive galaxies: $\alpha=-1.25$. This means that we either do not model low-mass galaxies very well, or that the mass-to-light ratio is not constant. The first option is probably true anyway since we cannot model merging of galaxies with masses below our lower limit towards the low-mass end of the mass function that we try to fit. With this in mind it is fair to say that the Schechter function does fit rather well for the limited set. We find $B_{\rm J}^*=-20.3$ for this set, which corresponds remarkably well with the value that Colless (1989) found for a sample of 14 observed clusters. It compares less well with $B_{\rm R}^*=-22.6$ found by Vink \& Katgert (1994) for a sample of 80 clusters, corresponding to $B_{\rm J}^*\approx-20.8$. Still we can say that the modelling performes reasonably well given the uncertainties in both the fits and the assumption of a constant mass-to-light ratio. \acknowledgments Joshua Barnes and Piet Hut are gratefully acknowledged for allowing use of their treecode, Edmund Bertschinger and Rien van de Weygaert for their code to generate initial conditions, and Eric Deul for allowing me to use the computer systems that are part of the DENIS project. I acknowledge EelcoSoft Software Services for partial financial support during the early stages of the project, and an European Community Research Rellowship as part of the HCM programme during its final stages.
\section{\intro}Introduction Recently the concept of time-symmetric elements of reality, introduced by Vaidman (1996a, 1997), stirred up a lively controversy which culminated in the joint publication of two papers in this journal (Kastner, 1999; Vaidman 1999). Using the standard formalism of standard quantum mechanics, one calculates the Born probability $$ P_B(a_i)=|\sandwich{\Psi}{{\bf P}_{A=a_i}}{\Psi}|,\eqno{(1)} $$ where the operator ${\bf P}_{A=a_i}$ projects on the subspace corresponding to the eigenvalue $a_i$ of the observable $A$. $P_B(a_i)$ is generally regarded as {\it the} probability with which a measurement of $A$ performed after the `preparation' of a system $S$ in the `state' $\ket{\Psi}$ yields the result $a_i$. But $P_B(a_i)$ is not the only such probability. Using a nonstandard formulation of standard quantum theory called time-symmetrized quantum theory (Aharonov and Vaidman, 1991; Vaidman, 1998), one calculates the ABL probability $$ P_{ABL}(a_i)={\absosq{ \sandwich{\Psi_2}{{\bf P}_{A=a_i}}{\Psi_1} }\over \Sigma_j\absosq{ \sandwich{\Psi_2}{{\bf P}_{A=a_j}}{\Psi_1} }}.\eqno{(2)} $$ ABL probabilities were first introduced in a seminal paper by Aharonov, Bergmann, and Lebowitz (1964). In this paper it was shown that $P_B(a_i)$ can also be thought of as the probability with which a measurement of $A$, performed {\it before} what may be called the `retroparation' of $S$ in the `state' $\ket{\Psi}$, yields the result $a_i$. Further it was shown that if a system is `prepared' at the time $t_1$ and `retropared' at the time $t_2$ in the respective `states' $\ket{\Psi_1}$ and $\ket{\Psi_2}$, the probability with which a measurement of $A$ performed at an intermediate time $t_m$ yields (or would have yielded) the result $a_i$, is given by $P_{ABL}(a_i)$.\fnote{% The $\Psi$'s in (2) are related to the `pre-/retropared' $\Psi$'s via unitary transformations\break $U(t_m-t_1)$ and $U(t_m-t_2)$.} Born probabilities can be measured (as relative frequencies) using preselected ensembles (that is, ensembles of identically `prepared' systems). ABL probabilities can be measured using pre- and postselected ensembles (that is, ensembles of systems that are both identically `prepared' and identically `retropared'). If the Born probability $P_B(a_i,t)$ of obtaining the result $a_i$ at time $t$ is equal to~1, one feels justified in regarding the value $a_i$ of the observable $A$ as a property that is actually possessed at the time $t$, that is, one feels justified in assuming that at the time $t$ there is is an {\it element of reality} corresponding to the value $a_i$ of the observable $A$ irrespective of whether $A$ is actually measured. Redhead (1987) has expressed this feeling as the following `sufficiency condition': \medskip{\leftskip=\parindent\rightskip=\parindent\noindent (ER1) If we can predict with certainty, or at any rate with [Born] probability one, the result of measuring a physical quantity at time $t$, then at the time $t$ there exists an element of reality corresponding to the physical quantity and having a value equal to the predicted measurement result.\par}\medskip The controversy about time-symmetric elements of reality arose because it appeared that Vaidman (1993) made the same claim with regard to ABL probabilities: \medskip{\leftskip=\parindent\rightskip=\parindent\noindent (ER2) If we can infer with certainty [that is, with ABL probability one] that the result of measuring at time $t$ an observable $A$ is $a$, then at the time $t$ there exists an element of reality $A=a$.\par}\medskip In response to criticism by Kastner and others (Kastner, 1999; and references therein), Vaidman (1999) clarified that he intended the term `element of reality' in a `technical' rather than `ontological' sense: saying that there is an element of reality $A=a$ is the same as saying that if $A$ is measured, the result is certain to be $a$. In other words, (ER2) is a tautology: if $A=a$ is certain to be found then $A=a$ is certain to be found. In formulating (ER2), Vaidman does not affirm the existence of an element of reality {\it irrespective} of whether $A$ is actually measured. (ER2) {\it defines} what it means to affirm the existence of an element of reality corresponding to $A=a$. To say that there is such an element of reality is to affirm the truth of a conditional, not the existence of an actual situation or state of affairs. Since ordinarily the locution `element of reality' refers to an actual state of affairs, Vaidman's terminological choice was unfortunate and has mislead many readers. But beyond that, his reading of (ER2) is unobjectionable. I shall, however, stick to the ordinary, ontological meaning of `element of reality.' In what follows, (ER2) is to be understood accordingly, that is, as affirming an actual state of affairs (the existence of an `ordinary' element of reality) just in case the corresponding ABL probability is one. Hence my showing that (ER2), thus understood, is false, has no bearing on Vaidman's reading of (ER2). A definition cannot be false. The aim of this paper is to show that not only (ER2) but also (ER1) is false. In Sec.~2 I discuss the three-box {\it gedanken} experiment due to Vaidman (1996b). By calculating the ABL probabilities associated with different versions of this experiment I show that positions are extrinsic, and that, consequently, (ER1) and (ER2) are both false. Since the validity of arguments based on time-symmetric quantum counterfactuals is open to debate, in Sec.~3 I show without making use of ABL probabilities that positions are extrinsic and that (ER1) is false. This leads to the conclusion that the measurement problem is a pseudoproblem, and that all that ever gets objectively entangled is counterfactuals. In Sec.~4 I establish the cogency of the argument of Sec.~2 by showing that the proper condition for the truth of a quantum counterfactual is $P_{ABL}=1$. This necessitates a discussion of objective probabilities, retroactive causality, and the objectivity or otherwise of the psychological arrow of time. In Sec.~5 I argue that since quantum mechanics presupposes the occurrence/existence of actual events and/or states of affairs, it cannot be called upon to account for the emergence of `classicality.' What is more, if quantum mechanics is as fundamental as its mathematical simplicity and empirical success suggest, the property-defining events or states of affairs presupposed by quantum mechanics are causal primaries -- nothing accounts for their occurrence or existence. If this is correct, the remaining interpretative task consists not in explaining the quantum-mechanical correlations and/or correlata but in understanding what they are trying to tell us about the world. I confine myself to pointing out, in Sec.~6, the most notable implications of the diachronic correlations, viz., the existence of entities of limited transtemporal identity, objective indefiniteness, and the spatial nonseparability of the world. The extrinsic nature of positions, finally, appears to involve a twofold vicious regress. Its resolution involves macroscopic objects, which are defined and discussed in Sec.~7. Section~8 concludes with a remark on the tension of contrast between objective indefiniteness and the inherent definiteness of language. \section{\vaidboxes}The Lesson of the Three-Box Experiment In the following I present a somewhat different but conceptually equivalent version of Vaidman's (1996b) three-box experiment. Consider a wall in which there are three holes $A$, $B$ and $C$. In front of the wall there is a particle source $Q$. Behind the wall there is a particle detector $D$. Both $Q$ and $D$ are equidistant from the three holes. Behind $C$ there is one other device; its purpose is to cause a phase shift by $\pi$. Particles emerging from the wall are thus preselected in a `state' $\ket{\Psi_1}$ proportional to $\ket{a}+\ket{b}+\ket{c}$, where $\ket{a}$, $\ket{b}$ and $\ket{c}$ represent the respective alternatives `particle goes through $A$,' `particle goes through $B$,' and `particle goes through $C$,' while detected particles are postselected in a `state' $\ket{\Psi_2}$ proportional to $\ket{a}+\ket{b}-\ket{c}$. We will consider two possible intermediate measurements. First we place near $A$ a device $F_a$ that beeps whenever a particle passes through $A$. With the help of the ABL formula one finds that every particle of this particular pre- and postselected ensemble $\cal E$ causes $F_a$ to beep with probability~1, as one may verify by calculating the probability with which a particle would be found passing through the union $B\cup C$ of $B$ and $C$: $$ P_{ABL}(\subset B\cup C)\propto \absosq{ \sandwich{\Psi_2}{{\bf P}_{\subset B\cup C}}{\Psi_1} }=0, $$ where ${\bf P}_{\subset B\cup C}=\ketbra bb+\ketbra cc$ projects on the subspace corresponding to the alternative `particle goes through $B\cup C$.' We obtain the same result by considering what would happen if $A$ were closed, or if all particles that make $F_a$ beep were removed from $\cal E$. The remaining particles are pre- and postselected in `states' proportional to $\ket{b}+\ket{c}$ and $\ket{b}-\ket{c}$, respectively, and these `states' are orthogonal. The result is an empty ensemble: if $A$ were closed, no particle would arrive at $D$. Does this warrant the conclusion that all particles belonging to $\cal E$ pass through $A$? Let us instead place near $B$ a device $F_b$ that beeps whenever a particle passes through $B$. Considering the invariance of $\ket{\Psi_1}$ and $\ket{\Psi_2}$ under interchange of $\ket{a}$ and $\ket{b}$, one is not surprised to find that the ensemble $\cal E$ would be empty if the particles causing $F_b$ to beep were removed. Hence if the conclusion that all particles belonging to $\cal E$ pass through $A$ is warranted, so is the conclusion that the same particles also pass through $B$. If these `conclusions' were legitimate, they would make nonsense of the very concept of localization. Therefore we are forced to conclude instead that an ABL probability equal to 1 does {\it not} warrant the existence of a corresponding element of reality (in the straightforward, ontological sense). Taken in this sense, (ER2) is false. It pays to investigate further. We are in fact dealing with four different experimental arrangements: (i)~there is no beeper, (ii)~$F_a$ is the only beeper in place, (iii)~$F_b$ is the only beeper in place, (iv)~both $F_a$ and $F_b$ are in place. The first arrangement permits no legitimate inference concerning the hole taken by a particle. Assuming that $F_a$ is 100\% efficient, the second arrangement guarantees that one of two inferences is warranted: `the particle goes through $A$' (in case $F_a$ beeps) or `the particle goes through $B\cup C$' (in case $F_a$ fails to beep). Assuming that $F_b$ is equally efficient, the third arrangement likewise guarantees that one of two inferences is warranted: `the particle goes through $B$' or `the particle goes through $A\cup C$.' The fourth arrangement, finally, guarantees that one of three inferences is warranted: `the particle goes through $A$' (in case $F_a$ beeps), `the particle goes through $B$' (in case $F_b$ beeps), and `the particle goes through $C$' (in case neither $F_a$ nor $F_b$ beeps). The following counterfactuals are therefore true: (i)~If $F_a$ were in place, either it would beep and the particle would go through $A$, or it would fail to beep and the particle would go through $B\cup C$. (ii)~If $F_b$ were in place, either it would beep and the particle would go through $B$, or it would fail to beep and the particle would go through $A\cup C$. (iii)~If both $F_a$ and $F_b$ were in place, one of the following three conjunctions would be true: $F_a$ beeps and the particle goes through $A$; $F_b$ beeps and the particle goes through $B$; neither beeper beeps and the particle goes through $C$. If we confine the discussion to particles that are emitted by $Q$ and detected by $D$, then the following counterfactuals are true: If $F_a$ but not $F_b$ were present, the alternatives represented by $\ket{b}$ and $\ket{c}$ would interfere with each other but not with the alternative represented by $\ket{a}$; as a consequence, they would interfere destructively; therefore $F_a$ would beep and the particle would go through $A$. By the same token, if $F_b$ were the only beeper present, the alternatives represented by $\ket{a}$ and $\ket{c}$ would interfere destructively, $F_b$ would beep, and the particle would go through $B$. Finally, if both beepers were present, no interference would take place; each particle would go through a particular hole, but not all particles would go through the same hole. To my mind, these counterfactuals are unobjectionable. One does not have to delve into the general philosophy of counterfactuals to see that they are true. I concur with Vaidman (1999) that quantum counterfactuals are unambiguous. Quantum counterfactuals are statements about possible worlds in which the outcomes of all measurements but one are the same as in the actual world. The remaining measurement is performed in a number of possible worlds (the number depends on the range of possible values) but not in the actual world. The three-box (or three-hole) experiment demonstrates that position probabilities cannot be assigned independently of experimental arrangements. More specifically, they cannot be assigned without specifying a set of experimentally distinguishable alternatives. A position probability of~1 depends not only on the way the particle is `prepared' and `retropared' but also on the set $L$ of alternative locations that can be experimentally distinguished. If $L=\{A,B\cup C\}$, the particle is certain to be found in (or going through) $A$, but the inference of an element of reality `the particle went through $A$' is warranted only if the members of $L$ are actually distinguished (that is, only if the corresponding experiment is actually performed). It follows that the position of a particle is an extrinsic property. By an {\it extrinsic} property $p$ of $S$ I mean a property of $S$ that is undefined unless either the truth or the falsity of the proposition ${\bf p}=$~`$S$~is~$p$' can be inferred from what happens or is the case in the `rest of the world' ${\cal W}-S$. The position of a particle is undefined unless there is a specific set $\{R_i\}$ of alternative locations, and unless there is a matter of fact about the particular location $R_j$ at which the particle is, or has been, present. (By `a matter of fact about the particular location $R_j$' I mean an actual event or an actual state of affairs from which that location can be inferred.\fnote{% Examples of actual events are the click of a Geiger counter or the deflection of a pointer needle. An actual state of affairs is expressed, for instance, by the statement `The needle points to the left.' Can such events and states of affairs be defined in quantum-mechanical terms? See below.}% ) Positions are {\it defined} in terms of position-indicating matters of fact. They `dangle' from actual events or actual states of affairs. And if it is true that `[t]here is nothing in quantum theory making it applicable to three atoms and inapplicable to $10^{23}$' (Peres and Zurek, 1982), this must be as true of footballs and cats as it is of particles and atoms. The positions of things {\it are} what matters of fact imply concerning the positions of things. If this is correct then (ER1) is as false as (ER2). In particular, the `sufficiency condition' (ER1) is {\it not} sufficient for the presence of a material object $O$ in a region of space $R$. The condition that is both necessary and sufficient for the presence of $O$ in $R$, is the existence of a matter of fact that indicates $O$'s presence in $R$. If there isn't any such matter of fact (now or anytime past or future), and if there also isn't any matter of fact that indicates $O$'s absence from $R$, then the sentence `$O$ is in $R$' is neither true nor false but meaningless, and $O$'s position with respect to $R$ (inside or outside) is undefined. \section{\pcer}Probabilities, Conditonals, Elements of Reality, and the Measurement Problem In the previous section I made use of the ABL probabilities associated with different versions of Vaidman's three-box experiment to show that positions are extrinsic, and that, consequently, both (ER1) and (ER2) are false. The validity of arguments based on time-symmetric quantum counterfactuals might be challenged. In the present section I therefore show without recourse to ABL probabilities that positions are extrinsic and that (ER1) is false. In the following section I shall establish the cogency of the argument of the previous section by showing that the proper condition for the truth of a quantum counterfactual is $P_{ABL}=1$. (It is readily verified that $P_B=1$ is sufficient but not necessary for this condition to be met.) Consider two perfect detectors $D_1$ and $D_2$ whose respective (disjoint) sensitive regions are $R_1$ and $R_2$. If the support of the (normalized) wave function associated with the (center-of-mass) position of an object $O$ is neither wholly inside $R_1$ nor wholly inside $R_2$, nothing necessitates the detection of $O$ by $D_1$, and nothing necessitates the detection of $O$ by $D_2$. But if the wave function vanishes outside $R_1\cup R_2$, the probabilities for either of the detectors to click add up to~1, so either of the detectors is certain to click. Two perfect detectors with sensitive regions $R_1$ and $R_2$ constitute one perfect detector $D$ with sensitive region $R_1\cup R_2$. But how can it be certain that one detector will click when individually neither detector is certain to click? What could cause $D$ to click while causing neither $D_1$ nor $D_2$ to click? That two perfect detectors with disjoint sensitive regions constitute one perfect detector for the union of the two regions forms part of the {\it definition} of what we mean by a perfect detector. By definition, a perfect detector clicks when the quantum-mechanical probability for it to click is~1. $D$ is certain to click because the probabilities for either of the two detectors to click add up to~1. Hence the question of what {\it causes} $D$ to click does not arise. Perfect detectors are theoretical constructs that by definition behave in a certain way. If real detectors would behave in the same way, it would be proper to inquire why they behave in this way. But real detectors are not perfect and do not behave in this way. A real detector is not certain to click when the corresponding quantum-mechanical probability is~1. Hence the question of what causes a real detector to click does not arise. Nothing causes a real detector to click. What I aim at in this paper is an interpretation of quantum mechanics that takes standard quantum mechanics to be fundamental and complete. My claim that nothing causes a real detector to click is based (i)~on this assumption and (ii)~on the observation that the efficiency of a real detector cannot be accounted for in quantum-mechanical terms. All quantum-mechanical probability assignments are relative to {\it perfect} detectors. If quantum mechanics predicts that $D_1$ will click with a probability of $1/2$, it does {\it not} predict that a {\it real} detector will click in 50\% of all runs of the actual experiment. What it predicts is that $D_1$ will click in 50\% of {\it those} runs of the experiment in which either $D_1$ or $D_2$ clicks. Quantum mechanics has nothing to say about the percentage of runs in which no counter clicks (that is, is tells us nothing about the efficiency of $D$, or of any other real detector for that matter). If quantum mechanics predicts that $D_1$ will click with probability~1, it accounts for the fact that whenever either $D_1$ or $D_2$ clicks, it is $D_1$ that clicks. It does {\it not} account for the clicking of either $D_1$ or $D_2$. Where quantum mechanics is concerned, nothing causes the clicking. And if quantum mechanics is as fundamental and complete as is here assumed, then this is true without qualification: nothing causes the clicking.\fnote{% It is well known that all actually existing detectors are less than perfect. On the other hand, there is no (obvious) theoretical limit to the efficiency of a real detector. It might one day be possible to build a detector with an efficiency arbitrarily close to 100\%. However, unless the efficiency of detectors is exactly 100\%, it remains impossible to interpret a `preparation' that warrants assigning probability~1 as {\it causing} a detector to click. If the preparation is to be a sufficient reason for the click, the detector must {\it always} click (that is, it must be perfect). What if it were possible to build perfect detectors? We could then speak of the preparation as the cause of the click, but if quantum mechanics is fundamental and complete, it would still be impossible to explain how the preparation causes the click: {\it ex hypothesi}, no underlying mechanism exists. The perfect correlation between preparation and click would have to be accepted as a brute fact. So would the fact that either $D_1$ or $D_2$ clicks when neither of them is certain to click. Causality would be just another name for such correlations, not an explanation.} Quantum mechanics assigns probabilities (whether Born or ABL) to alternative events (e.g., deflection of the pointer needle to the left or to the right) or to alternative states of affairs (e.g., the needle's pointing left or right). Implicit in every normalized distribution of probabilities over a specified set of alternative events or states of affairs is the assumption that exactly one of the specified alternatives happens or obtains. If we assign normalized probabilities to a set of counterfactuals, we still assume (counterfactually) that exactly one of the counterfactuals is true. In other words, if we assign probabilities to the possible results of an unperformed measurement, we still assume that the measurement, if it had been performed, would have yielded a definite result. Like all (normalized) probabilities, the probabilities assigned by quantum mechanics are assigned to mutually exclusive and jointly exhaustive possibilities, and they are assigned {\it on the supposition} that exactly one possibility is, or would have been, a fact. Even the predictions of the standard version of standard quantum mechanics therefore are {\it conditionals}. Everything this version tells us conforms to the following pattern: {\it If} there is going to be a matter of fact about the alternative taken (from a specific range of alternatives), {\it then} such and such are the Born probabilities with which that matter of fact will indicate this or that alternative. It is important to understand that quantum mechanics never allows us to predict {\it that} there will be such a matter of fact, unconditionally. If the Born probability of a particular event $F$ is~1, we are not entitled to predict that $F$ will happen. What we are entitled to infer is only this: Given that one of a specified set of events will happen, and given that $F$ is an element of this set, the event that will happen is $F$. In order to get from a true conditional to an element of reality, a condition has to be met: a measurement must be successfully performed, there must be a matter of fact about the value of an observable, one of a specific set of alternative property-indicating events or states of affairs must happen or obtain. Quantum mechanics does not predict that a measurement will take place, nor the time at which one will take place, nor does it specify the conditions in which one will take place. It requires us to {\it assume} that one will take place, for it is on this assumption that its probability assignments are founded. It follows that (ER1) is false. A Born probability equal to~1 is equivalent to a conditional $c$. The inference of a corresponding element of reality is warranted only if the condition laid down by $c$ is actually met. It also follows that positions are extrinsic. The condition laid down by $c$ is the existence of a matter of fact about the value taken by some observable. If this observable has for its spectrum a set $\{R_i\}$ of mutually disjoint regions of space, if $R$ is an element of $\{R_i\}$, and if the Born probability associated with $R$ is~1, then $O$ is inside $R$ just in case there is a matter of fact about the particular element of $\{R_i\}$ that contains $O$. I conclude this section with a few remarks concerning the so-called measurement problem. First some basic facts. Quantum mechanics represents the possible values $q^k_i$ of all observables $Q^k$ as projection operators ${\bf P}_{Q^k=q^k_i}$ on some Hilbert space $\cal H$. The projection operators that jointly represent the range of possible values of a given observable are mutually orthogonal. If one defines the `state' of a system as a {\it probability measure} on the projection operators on $\cal H$ (Cassinello and S\'anchez-G\'omez, 1996; Jauch, 1968, p. 94) resulting from a {\it preparation} of the system (Jauch, 1968, p. 92), one finds (Cassinello and S\'anchez-G\'omez, 1996; Jauch, 1968, p. 132) that every such probability measure has the form $P({\bf P})=\hbox{Tr}({\bf WP})$, where $\bf W$ is a unique density operator (that is, a unique self-adjoint, positive operator satisfying $\hbox{Tr}({\bf W})=1$ and ${\bf W}^2\leq{\bf W}$). [The trace $\hbox{Tr}({\bf X})$ is the sum $\sum_i\sandwich{i}{{\bf X}}{i}$, where $\{\ket{i}\}$ is any orthonormal basis in $\cal H$.] If ${\bf W}^2(t)={\bf W}(t)$, ${\bf W}(t)$ projects on a one-dimensional subspace of $\cal H$ and thus is equivalent -- apart from an irrelevant phase factor -- to a `state' vector $\ket{\Psi(t)}$ or a wave function $\Psi(x,t)$, $x$ being any point in the system's configuration space. In this case one retrieves the Born formula~(1). The quantum-mechanical `state' vector (or the wave function, or the density operator) thus is essentially a probability measure on the projection operators on $\cal H$, specifying probability distributions over all sets of mutually orthogonal subspaces of $\cal H$. Hence the $t$ that appears in the `states' ${\bf W}(t)$, $\ket{\Psi(t)}$, and $\Psi(x,t)$ has the same significance as the $t$ that appears in the time-dependent probabilities $P_B(q^k_i,t)$. Now recall that quantum mechanics predicts neither that a measurement will take place nor the time at which one will take place. It requires us to {\it assume} that a measurement will take place {\it at a specified time}. The time-dependence of the `state' vector therefore is a dependence on the {\it specified} time at which a {\it specified} observable (with a {\it specified} range of values) is measured either in the actual world or in a set of possible worlds. It is {\it not} the time-dependence of a state of affairs that evolves in time. On the supposition that $\ket{\Psi(t)}$ represents a state of affairs that evolves in time (so that at every time $t$ a state of affairs $\ket{\Psi(t)}$ obtains), one needs to explain what brings about the real or apparent discontinuous transition from a state of affairs represented by a ket of the form $\ket{\Psi(t)}=\sum_i a_i(t)\ket{a_i}$ to the state of affairs represented by one of the kets $\ket{a_i}$. This is the measurement problem. It is a pseudoproblem because a collection of time-dependent probabilities is not a state of affairs that evolves in time. The probability for something to happen at the time $t$ does not exist at $t$, any more than the probability for something to be located in $R$ exists in $R$. The probabilities $P_B(q^k_i,t)$ are determined by the relevant matters of fact about the properties possessed by a physical system at or before a certain time $t_0$. In the special case of a complete measurement performed at $t_0$, they are given by the Born formula $$ P_B(q^k_i,t)=|\sandwich{\Psi(t)}{{\bf P}_{Q^k=q^k_i}}{\Psi(t)}|\qquad \hbox{for $t\geq t_0$},\eqno{(3)} $$ where $\ket{\Psi(t)}=U(t-t_0)\ket{\Psi(t_0)}$. $\ket{\Psi(t_0)}$ is the `state' `prepared' by the measurement at $t_0$ (that is, it represents the properties possessed by the system at $t_0$). $U(t-t_0)$ is the unitary operator that governs the time-dependence of quantum-mechanical probabilities (often misleadingly referred to as the `time evolution operator'). And $t$ is the {\it stipulated} time at which the next measurement is performed, either actually or counterfactually. Thus all that a superposition of the form $\ket{\Psi(t)}=\sum_i a_i(t)\ket{a_i}$ tells us, is this: {\it If} there is a matter of fact from which one can infer the particular property (from the set of properties represented by the kets $\ket{a_i}$) that is actually possessed by the system at the stipulated time $t$, then the prior probability that this matter of fact indicates the property represented by $\ket{a_i}$ is $\absosq{a_i}$. It is self-evident that if there {\it is} such a matter of fact, and if this matter of fact is taken into account, the correct basis for further conditional inferences is not $\ket{\Psi(t)}$ but one of the kets $\ket{a_i}$. This obvious truism is the entire content of the so-called projection postulate (L\"uders, 1951; von Neumann, 1955). By the same token, all that an entangled `state' of the form $\sum_i a_i(t)\ket{b_i}\otimes\ket{a_i}$ tells us, is this: {\it If} there are two matters of fact, one indicating which of the properties represented by the kets $\ket{a_i}$ is possessed by the first system, and another indicating which of the properties represented by the kets $\ket{b_i}$ is possessed by the second system, then the two matters of fact together indicate $\ket{a_i}$ and $\ket{b_i}$ with probability $\absosq{a_i}$, and they indicate $\ket{a_i}$ and $\ket{b_j}$ ($j\neq i$) with probability~0. But if there {\it is} any such matter of fact, these entangled probabilities are based on an incomplete set of facts and are therefore subjective (that is, they reflect our ignorance of some relevant fact). All that ever gets {\it objectively} entangled is {\it counterfactuals}. \section{\opretro}Objective Probabilities, Retrocausation, and the Arrow of Time What most strikingly distinguishes quantum physics from classical physics is the existence of objective probabilities. In a classical world there are no (nontrivial) objective probabilities: the probability of dealing an ace is not $1/13$ but either $1$ or $0$, depending on whether or not an ace is top card. Objective probabilities have nothing to do with ignorance; there is nothing (that is, no actual state of affairs, no actually possessed property, no actually obtained measurement result) for us to be ignorant of. Then what is it that {\it has} an objective probability? What are objective probabilities distributed over? The obvious answer is: counterfactuals. Only a contrary-to-fact conditional can be assigned an objective probability. Objective probabilities are distributed over the possible results of {\it unperformed} measurements. Objective probabilities are objective in the sense that they are not subjective, and they are not subjective because they would be so only if the corresponding measurements were performed. In short, objective probabilities are probabilities that are {\it counterfactually subjective}. Probabilities can be objective only if they are based on a complete set of facts. Otherwise they are subjective: they reflect our ignorance of some of the relevant facts. Born probabilities are in general calculated on the basis of an incomplete set of facts; they take into account the relevant past matters of fact but ignore the relevant future matters of fact. Born probabilities are objective only if there are no relevant future matters of fact. Thus they cannot be objective if any one of the measurements to the possible results of which they are assigned, is actually performed. This is equally true of ABL probabilities: if one of the measurements to the possible results of which they are assigned, is actually performed, they too are calculated on the basis of an incomplete set of facts. They take into account all revelant matters of fact except the result of the actually performed measurement. On the other hand, if none of these measurements is actually performed, ABL probabilities take into account all relevant matters of fact and are therefore objective. Thus probabilities are objective only if they are distributed over alternative properties or values none of which are actually possessed, and only if they are based on all relevant events or states of affairs, including those that are yet to occur or obtain. In general the objective probabilities associated with contrary-to-fact conditionals depend also on events that haven't yet happened or states of affairs that are yet to obtain. Hence some kind of retroactive causality appears to be at work. This necessitates a few remarks concerning causality and the apparent `flow' of time. But first let us note that nothing entails the existence of time-reversed causal connections between {\it actual} events and/or states of affairs. To take a concrete example, suppose that at $t_1$ the $x$ component $\sigma_x$ of the spin of an electron is measured, that at $t_2>t_1$ $\sigma_y$ is measured, and that the respective results are $\uparrow_x$ and $\uparrow_y$. Then a measurement of $\sigma_x$ would have yielded $\uparrow_x$ if it had been performed at an intermediate time $t_m$, and a measurement of $\sigma_y$ would have yielded $\uparrow_y$ if it had been performed instead. What would have happened at $t_m$ depends not only on what happens at $t_1$ but also on what happens at $t_2$. But if either $\sigma_x$ or $\sigma_y$ is actually measured at $t_m$ (other things being equal), nothing compels us to take the view that $\uparrow_y$ was obtained at $t_m$ {\it because} the same result was obtained at $t_2$. We can stick to the idea that causes precede their effects, according to which $\uparrow_y$ was obtained at $t_2$ {\it because} the same result was obtained at $t_m$. The point, however, is that nothing in the physics {\it prevents} us from taking the opposite view. The distinction we make between a cause and its effect is based on the apparent `motion' of our location in time -- the present moment -- toward the future. This special location and its apparent `motion' are as extraneous to physics as are our location and motion in space (Price, 1996). Equally extraneous, therefore, is the distinction between causes and effects. Physics deals with correlations between actual events or states of affairs, classical physics with deterministic correlations, quantum physics with statistical ones. Classical physics allows us to explain the deterministic correlations (abstracted from what appear to be universal regularities) in terms of causal links between individual events. And for some reason to be explained presently, we identify the earlier of two diachronically correlated events as the cause and the later as the effect. The time symmetry of the classical laws of motion, however, makes it equally possible to take the opposite view, according to which the later event is the cause and the earlier event the effect. In a deterministic world, the state of affairs at any time $t$ determines the state of affairs at any other time $t'$, irrespective of the temporal order of $t$ and $t'$. The belief in a time-asymmetric {\it physical} causality is nothing but an animistic projection of the perspective of a conscious agent into the inanimate world, as I proceed to show. I conceive of myself as a causal agent with a certain freedom of choice. But I cannot conceive of my choice as exerting a causal influence on anything that I knew, or could have known, at the time $t_c$ of my choice. I can conceive of my choice as causally determining only such events or states of affairs as are unknowable to me at $t_c$. On the simplest account, what I knew or could have known at $t_c$ is everything that happened before $t_c$. And what is unknowable to me at $t_c$ is everything that will happen thereafter. This is the reason why we tend to believe that we can causally influence the future but not the past. And this constraint on {\it our} (real or imagined) causal efficacy is what we impose, without justification, both on the deterministic world of classical physics and on the indeterministic world of quantum physics. In my goal-directed activities I exploit the time-symmetric laws of physics. When I kick a football with the intention of scoring a goal, I make (implicit or explicit) use of my knowledge of the time-symmetric law that governs the ball's trajectory. But my thinking of the kick as the cause and of the scored goal as its effect has nothing to do with the underlying physics. It has everything to do with my self-perception as an agent and my successive experience of the world. The time-asymmetric causality of a conscious agent in a successively experienced world rides piggyback on the symmetric determinisms of the physical world, and in general it rides into the future because in general the future is what is unknowable to us. But it may also ride into the past. Three factors account for this possibility. First, as I said, the underlying physics is time-symmetric. If we ignore the strange case of the neutral kaon (which doesn't appear to be relevant to the interpretation of quantum mechanics), this is as true of quantum physics as it is of classical physics. If the standard formulation of quantum physics is asymmetric with respect to time, it is because we think (again without justification) that a measurement does more than yield a particular result. We tend to think that it also {\it prepares} a state of affairs which evolves toward the future. But if this is a consistent way of thinking -- it is {\it not} (Mohrhoff, 1999) -- then it is equally consistent to think that a measurement `retropares' a state of affairs that evolves toward the past, as Aharonov, Bergmann, and Lebowitz (1964) have shown. Second, what matters is what can be known. If I could know the future, I could not conceive of it as causally dependent on my present choice. In fact, if I could (in principle) know both the past and the future, I could not conceive of myself as an agent. I can conceive of my choice as causally determining the future precisely because I cannot know the future. This has nothing to do with the truism that the future does not (yet) exist. Even if the future in some way `already' exists, it can in part be determined by my present choice, provided I cannot know it at the time of my choice. By the same token, a past state of affairs can be determined by my present choice, provided I cannot know that state of affairs before the choice is made. There are two possible reasons why a state of affairs $F$ cannot be known to me at a given time $t$: (i)~$F$ may obtain only after $t$; (ii)~at $t$ there may as yet exist no matter of fact from which $F$ can be inferred. This takes us to the last of the three factors which account for the possibility of retrocausation: the contingent properties of physical systems are extrinsic. By a {\it contingent} property I mean a property that may or may not be possessed by a given system at a given time. For example, being inside a given region of space and having a spin component of $+\hbar/2$ along a given axis are contingent properties of electrons. Properties that can be retrocausally determined by the choice of an experimenter, cannot be intrinsic. [A property $p$ of a physical system $S$ is {\it intrinsic} iff the proposition ${\bf p}=$~`$S$~is~$p$' is `of itself' (that is, unconditionally) either true or false at any time.] If $p$ is an extrinsic property of $S$, the respective criteria for the truth and the falsity of the proposition ${\bf p}=$~`$S$~is~$p$' are to be sought in the `rest of the world' ${\cal W}-S$, and it is possible that neither criterion is satisfied, in which case $\bf p$ is neither true nor false but meaningless. It is also possible that each criterion consists in an event that may occur only after the time to which $\bf p$ refers. If this event is to some extent determined by an experimenter's choice, retrocausation is at work. On the other hand, if $p$ is an intrinsic property of $S$, $\bf p$ has a truth value (`true' or `false') independently of what happens in ${\cal W}-S$, so {\it a fortiori} it has a truth value independently of what happens there after the time $t$ to which $\bf p$ refers. There is then no reason why the truth value of $\bf p$ should be unknowable until some time $t'>t$. In principle it is knowable at $t$, and therefore we cannot (or at any rate, need not) conceive of it as being to some extent determined by the experimenter's choice at $t'$. A paradigm case of retrocausation at work (Mohrhoff, 1999) is the experiment of Englert, Scully, and Walther (1994; Scully, Englert, and Walther, 1991). This experiment permits the experimenters to choose between (i)~measuring the phase relation with which a given atom emerges coherently from (the union of) two slits and (ii)~determining the particular slit from which the atom emerges. The experimenters can exert this choice after the atom has emerged from the slit plate and even after it has hit the screen. By choosing to create a matter of fact about the slit taken by the atom, they retroactively cause the atom to have passed through a particular slit. By choosing instead to create a matter of fact about the atom's phase relation, they retroactively cause the atom to have emerged with a definite phase relation. The retrocausal efficacy of their choice rests on the three factors listed above (in different order): (i)~The four propositions ${\bf a}_1=$~``the atom went through the first slit,'' ${\bf a}_2=$~``the atom went through the second slit,'' ${\bf a}_+=$~``the atom emerged from the slits in phase,'' and ${\bf a}_-=$~``the atom emerged from the slits out of phase'' affirm {\it extrinsic} properties. (ii)~There exist {\it time-symmetric} correlations between the atom's possible properties at the time of its passing the slit plate and the possible results of two mutually exclusive experiments that can be performed at a later time. (iii)~The result of the actually performed experiment is the first (earliest) matter of fact about either the particular slit taken by the atom or the phase relation with which the atom emerged from the slits. Before they made their choice, the experimenters could not possibly have {\it known} the slit from which, or the phase relation with which, the atom emerged. Probabilities, I said, can be objective only if they are based on all relevant matters of fact, including those still in the future. We are now in a position to see clearly why it should be so. Our distinction between the past, the present, and the future has nothing to do with physics. Physics knows nothing of the experiential {\it now} (the special moment at which the world has the technicolor reality it has in consciousness), nor does it know anything of the difference between what happened before {\it now} (the past) and what will happen after {\it now} (the future). \longsection{\qmcc}The World According to Quantum Mechanics: Fundamentally Inexplicable Correlations Between Fundamentally Inexplicable Events It is commonly believed that it is the business of quantum mechanics to account for the occurrence/existence of actual events or states of affairs. Environment-induced superselection (Joos and Zeh, 1985; Zurek, 1981, 1982), decoherent histories (Gell-Mann and Hartle, 1990; Griffiths, 1984; Omn\`es, 1992), quantum state diffusion (Gisin and Percival, 1992; Percival 1994), and spontaneous collapse (Ghirardi, Rimini and Weber, 1986; Pearle, 1989) are just some of the strategies that have been adopted with a view to explaining the emergence of `classicality.' Whatever is achieved by these interesting endeavors, they miss this crucial point: quantum mechanics only takes us from facts to probabilities of possible facts. The question of how it is that exactly one possibility is realized must not be asked of a formalism that serves to assign probabilities on the implicit {\it assumption} that exactly one of a specified set of possibilities {\it is} realized. Even the step from probability~1 to factuality crosses a gulf that quantum mechanics cannot bridge. Quantum mechanics can tell us that $O$ is certain to be found in $R$ {\it given that} there is a matter of fact about its presence or otherwise in $R$, but only the actual matter of fact warrants the inference that $O$ is in $R$. Quantum mechanics does not predict that a measurement will take place, nor the time at which one will take place, nor does it specify the conditions in which one will take place. And if quantum mechanics is as fundamental as I presume it is, {\it nothing} allows us to predict that or when a measurement takes place, or to specify conditions in which one is certain to take place, for there is nothing that {\it causes} a measurement to take place. In other words, a matter of fact about the value of an observable is a causal primary. A {\it causal primary} is an event or state of affairs the occurrence or existence of which is not necessitated by any cause, antecedent or otherwise. I do not mean to say that in general nothing causes a measurement to yield this rather than that particular value. Unless one postulates hidden variables, this is a triviality. What I mean to say is that nothing ever causes a measurement to take place. Measurements (and in clear this means detection events) are causal primaries. No detector is 100\% efficient. Using similar detectors in series, it is easy enough to experimentally establish a detector's (approximate) likelihood to click when the corresponding Born probability is~1, but of this likelihood no theoretical account is possible.\fnote{% There are two kinds of probability, the probability {\it that} a detector will respond (rather than not respond), and the probability that this (rather than any other) detector will respond {\it given} that exactly one detector will respond. The former probability cannot be calculated using the quantum formalism (nor, if quantum mechanics is fundamental and complete, any other formalism). One can of course analyze the efficiency of, say, a Geiger counter into the efficiencies of its `component detectors' (the ionization cross sections of the ionizable targets it contains), but the efficiencies of the `elementary detectors' cannot be analyzed any further. This entails that a fundamental coupling constant such as the fine structure constant cannot be calculated from `first principles;' it can only be gleaned from the experimental data.} {\it A fortiori}, no theoretical account is possible of why or when a detector is certain to click. It never is. Quantum physics thus is concerned with correlations between events or states of affairs that are uncaused and therefore fundamentally inexplicable. As physicists we are not likely to take kindly to this conclusion, which may account for the blind spot behind which its inevitability has been hidden so long. But we certainly {\it are} at a loss when it comes to accounting for the world of definite occurrences. Recently Mermin (1998) advocated an interpretation of the formalism of standard quantum mechanics according to which ``[c]orrelations have physical reality; that which they correlate, does not.'' He does not claim that there are no correlata, only that they are not part of {\it physical} reality. The correlated events belong to a larger reality which includes consciousness and which lies outside the scope of physics. Thus Mermin agrees that, where physics is concerned, the correlata {\it are} fundamentally inexplicable. The idea that the correlata are conscious perceptions (Lockwood, 1989; Page, 1996), or beliefs (Albert, 1992), or knowings (Stapp, 1993) has a respectable pedigree (London and Bauer, 1939; von Neumann, 1955). If one thinks of the state vector as representing a state of affairs that evolves in time, one needs something that is `more actual' than the state vector -- something that bestows `a higher degree of actuality' than does the state vector -- to explain why every successful measurement has exactly one result, or why measurements are possible at all. This is the spurious measurement problem all over again. It is spurious because the state vector does {\it not} represent an evolving state of affairs. If we were to relinquish this unwarranted notion, we would not need two kinds of reality to make sense of quantum mechanics, such as a physical reality and a reality that includes consciousness (Mermin, 1998), or a potential reality and an actual reality (Heisenberg, 1958; Popper, 1982; Shimony, 1978, 1989), or a mind-constructed `empirical' reality and a mind-independent `veiled' reality (d'Espagnat, 1995), or an unrecorded `smoky dragon' reality and an irreversibly recorded reality (Wheeler, 1983). We could confine ourselves to talking about events that are causal primaries, the inferences that are warranted by such events, the correlations between such events or such inferences, and the further inferences that are warranted by these correlations. I do not deny that there is a larger reality that includes consciousness and that lies outside the scope of physics. What I maintain is that the interpretative problems concerning quantum mechanics can be solved without appealing to any larger reality, and that such an appeal does not help solving those problems because it is neither necessary nor possible to account for the occurrence of a causal primary. Theoretical physics is partly mathematics and partly semantics. The semantic task is to name the fundamental epistemological and/or ontological entities and/or relations represented by the symbols of the formalism. I cannot think of a more satisfactory choice of a basic (and therefore not further explicable) ontological entity for a physical theory than a causal primary -- something that is inexplicable {\it by definition}. Ever since the seminal paper by Einstein, Podolsky, and Rosen (1935), it has been argued that quantum mechanics is incomplete (Bell, 1966; Ford and Mantica, 1992; Lockwood, 1989; Primas, 1990). In point of fact, no theory can be more complete (with regard to its subject matter) than one that accounts for everything (within its subject matter) but what is inexplicable by definition. If there is anything that is incomplete, it is reality itself (that is, reality is incomplete relative to our {\it description} of it, which is `overcomplete') -- but I'm getting ahead of myself. Because the occurrence/existence of actual events or states of affairs is presupposed by the formalism, locutions such as `actual event,' `actual state of affairs,' `matter of fact' cannot even be {\it defined} within the formalism. This conclusion too is unlikely to be popular with physicists, who naturally prefer to define their concepts in terms of the mathematical formalism they use. Einstein spent the last thirty years of his life trying (in vain) to get rid of field sources -- those entities that have the insolence to be real by themselves rather than by courtesy of some equation (Pais, 1982). Small wonder if he resisted Bohr's insight that not even the {\it properties} of things can be defined in purely mathematical terms. But Bohr was right. If Bohr (1934, 1963) insisted on the necessity of describing quantum phenomena in terms of experimental arrangements, it was because he held that the properties of quantum systems are {\it defined} by the experimental arrangements in which they are displayed (d'Espagnat, 1976). For `experimental arrangement' read: what matters of fact permit us to infer concerning the properties of a given system at a given time. The contingent properties of physical systems are defined in terms of the actual events or states of affairs from which they can be inferred. They `dangle' from what happens or is the case in the rest of the world. They cannot be defined in purely mathematical terms, for only intrinsic properties can be so defined. The scope of physics is not restricted to laboratory experiments. {\it Any} matter of fact that has a bearing on the properties of a physical system qualifies as a `measurement result.' What is relevant is the occurrence or existence of an event or state of affairs warranting the assertability of a statement of the form `$S$ is $p$ (at the time $t$),' irrespective of whether anyone is around to assert, or take cognizance, of that event or state of affairs, and irrespective of whether it has been anyone's intention to learn something about $S$. The following picture emerges. The world is a mass of events that are causal primaries. Without any correlations between these events, it would be a total chaos. As it turns out, the uncaused events are strongly correlated. If we don't look too closely, they fall into neat patterns that admit of being thought of as persistent objects with definite and continuously evolving positions. Projecting our time-asymmetric agent-causality into the time-symmetric world of physics, we think of the positions possessed at later times as causally determined by the positions possessed at earlier times. If we look more closely, we find that positions aren't always attributable, and that those that are attributable aren't always predictable on the basis of past events. We discover that positions do not `dangle' from earlier positions by causal strings but instead `dangle' from position-defining events that are statistically correlated but (being causal primaries) are not causally connected. Quantum mechanics describes the correlations but does nothing to explain them. Not only the correlata but also the correlations are incapable of (causal) explanation. Causal explanations are confined to the familiar macroworld of deterministic processes and things that evolve in time. This macroworld with its causal links is something we project onto the correlations and their uncaused correlata, but the projection works only to the extent that the correlations are not manifestly probabilistic.\fnote{% This is discussed in the last two sections.} There are no causal processes more fundamental than the correlations and their correlata, processes that could in any manner account for the correlations or the correlata. \section{\uinex}Spatial Nonseparability The remaining interpretative task thus consists not in explaining the correlations but in understanding what they are trying to tell us about the world. Here I will confine myself to discussing some of the implications of the diachronic correlations (the correlations between results of measurements performed on the same system at different times).\fnote{% The implications of the synchronic (EPR) correlations have been discussed elsewhere (Mohrhoff, submitted).} Perhaps the first insight one gleans from the correlations is the existence of persisting {\it entities}. If the correlations did not permit us to speak of such entities, we could not think of the correlata as possessed properties, extrinsic or otherwise. Suppose that we perform a series of position measurements. And suppose that every position measurement yields exactly one result (that is, each time exactly one detector clicks). Then we are entitled to infer the existence of an entity $O$ which persists through time (if not for all time), to think of the clicks given off by the detectors as matters of fact about the successive positions of this entity, to think of the behavior of the detectors as position measurements, and to think of the detectors as detectors. (The lack of transtemporal identity among particles of the same type of course forbids us to extend to such particles the individuality of a fully `classical' entity.) The successive positions of $O$, however, are extrinsic: they are what can be inferred from the pattern of clicks. All that can be inferred concerning $O$'s positions at times at which no detector clicks, is counterfactual and probabilistic.\fnote{% The detectors of the present scenario are assumed to be time-specific: a click not only indicates a position but also the time at which it is possessed.} There is a persistent entity all right, but there is then no actually possessed position to go with it. The next lesson to be learned from the correlations is that the positions of things are objectively indefinite or `fuzzy.' This does not mean that $O$ {\it has} as fuzzy position. It means that statements of the form `$O$ is in $R$ at $t$' are sometimes neither true nor false but meaningless. This possibility stands or falls with the extrinsic nature of positions and the existence of objective probabilities. Take the counterfactual `If there were a matter of fact about the slit taken by the atom, the atom would have taken the first slit.' We can assign to this counterfactual an {\it objective} probability iff the proposition `The atom went through the first slit' is neither true nor false but meaningless. The reason why this proposition {\it can} be meaningless is that positions are extrinsic. It {\it is} meaningless just in case there isn't any matter of fact about the slit taken by the atom. If it is true that the atom went through the union of the slits (that is, if the atom was emitted on one side of the slit plate and detected on the other side), and if it is meaningless to say that the atom went through the first slit (in which case it is also meaningless to say that it went through the second slit), then the conceptual distinction we make between the two slits has no reality for the atom. If that distinction were real for the atom (that is, if the atom behaved as if the two slits were distinct), the atom could not behave as if it went -- as a whole, without being divided into distinct parts -- simultaneously through both slits. But (if quantum mechanics is fundamental and complete) this is what the atom does when interference fringes are observed. Thus there are objects for which our conceptual distinction between mutually disjoint regions of space does not exist. It follows that the distinction between such regions cannot be real {\it per se} (that is, it cannot be an intrinsic property of the world). If it were real {\it per se}, the following would be the true: at any one time, for every finite region $R$, the world can be divided into things or parts that are situated inside $R$, and things or parts that are situated inside the complement $R'$ of $R$. The boundary of $R$ would demarcate an intrinsically distinct part of the world. But if this were the case, exactly one of the following three propositions would be true of every object $O$ at any given time: (i)~$O$ is situated wholly inside $R$; (ii)~$O$ is situated wholly inside $R'$; (iii)~$O$ has two parts, one situated wholly inside $R$ and one situated wholly inside $R'$. If there is anything that (standard) quantum mechanics is trying to tells us about the world, it is that for at least some objects all of these propositions are sometimes false. It follows that the multiplicity and the distinctions inherent in our {\it mathematical concept} of space -- a transfinite set of triplets of real numbers -- are {\it not} intrinsic features of {\it physical} space. The notion that these features of our mathematical concept of space are intrinsic to physical space -- in other words, the notion that the world is {\it spatially separable} -- is a delusion. This notion is as inconsistent with quantum mechanics as the notion of absolute simultaneity is with special relativity. `Here' and `there' are not {\it per se} distinct. Reality is {\it fundamentally nonseparable}. Like the positions of things, spatial distinctions `dangle' from actual events or states of affairs. Reality is not built on a space that is differentiated the way our mathematical concept of space is differentiated. A description of the world that incorporates such a space -- and {\it a fortiori} every description that identifies `the points of space (or space-time)' as the carriers of physical properties -- is `overcomplete.' Reality is built on matters of fact, and the actually existing differences between `here' and `there' are the differences that can be inferred from matters of fact. In and of itself, physical space -- or the reality underlying it -- is undifferentiated, one. \section{\mapicos}Macroscopic Objects The extrinsic nature of positions appears to involve a twofold vicious regress. To adequately deal with it, I need to talk about macroscopic objects. A {\it macroscopic object} $M$ is an object that satisfies the following criterion: any factually warranted inference concerning the position of $M$ at any time $t$ is predictable (with certainty) on the basis of factually warranted inferences about the positions of $M$ at earlier times. (A factually warranted inference is an inference that is warranted by some matter of fact.) Thus, to the extent that they can be inferred from actual events, the successive positions of a macroscopic object evolve deterministically. This makes it possible to ignore the fact that the positions of macroscopic objects, like all actually possessed positions, depend for their existence on position-indicating events. We can treat the positions of macroscopic objects as intrinsic properties and assume that they follow definite and causally determined trajectories, without ever risking to be contradicted by an actual event. I do not mean to say that the position of $M$ really {\it is} definite. Even the positions of macroscopic objects are fuzzy, albeit not manifestly so: the positional indefiniteness of $M$ does not evince itself through unpredictable position-indicating events. Nor do I mean to say that the positions of macroscopic objects really {\it are} intrinsic. They too `dangle' from actual events. But they do so in a way that is predictable, that does not reveal any fuzziness. We may think of macroscopic objects as following definite trajectories, or we may think of them as following fuzzy trajectories. Since all matters of fact about their positions are predictable, it makes no difference: the fuzziness has no factual consequences. Classical behavior results when the factually warranted positions fuse into a not manifestly fuzzy trajectory. It has little to do with the `classical limit' in which the wave packet shrinks to a continuously moving point, for the wave packet (of whatever size) is a bundle of probabilities associated with time-dependent counterfactuals, not the actual trajectory of an object.\fnote{% Good examples of how not to get from quantum to classical are the unsuccessful attempts to obtain the exponential decay law, which pertains to factually warranted inferences and is consistent with all experimental data, from the Schr\"odinger equation, which tells us how the probabilities associated with counterfactuals depend on time (Onley and Kumar, 1992; Singh and Whitaker, 1982).} By saying that matters of fact about the positions of macroscopic objects are predictable I do not mean that the {\it existence} of such a matter of fact is predictable. Once again, a Born probability equal to~1 does not warrant the prediction {\it that} an event will happen or {\it that} a state of affairs will obtain. Only if it is {\it taken for granted} that exactly one of a range of possible events or states of affairs will happen or obtain, does a Born probability equal to~1 allow us to predict {\it which} event or state of affairs will happen or obtain. What I mean by saying that matters of fact about the (successive) positions of a macroscopic object are predictable, is this: what an actual event or state of affairs implies regarding the position of a macroscopic object is consistent with what can be predicted with the help of some classical dynamical law on the basis of earlier position-defining events. Everything a macroscopic object does (that is, every matter of fact about its present properties) follows via the pertinent classical laws from what it did (that is, from matters of fact about its past properties).\fnote{% The above definition of `macroscopic' does not stipulate that events indicating departures from the classically predicted behavior occur with zero {\it probability}. An object is entitled to the label `macroscopic' if no such event {\it actually} occurs during its lifetime. What matters is not whether such an event {\it may} occur (with whatever probability) but whether it ever {\it does} occur.} When I speak of the {\it existence} of a matter of fact, I mean the occurrence of an actual event or the existence of an actual state of affairs. It is worth emphasizing that this is something that cannot be undone or `erased' (Englert, Scully, and Walther, 1999; Mohrhoff, 1999). According to Wheeler's interpretation of the Copenhagen interpretation, `no elementary quantum phenomenon is a phenomenon until it is registered, recorded, ``brought to a close'' by an ``irreversible act of amplification,'' such as the blackening of a grain of photographic emulsion or the triggering of a counter' (Wheeler, 1983). In point of fact, there is no such thing as an `irreversible act of amplification.' As long as what is `amplified' is counterfactuals, the `act of amplification' is reversible. No amount of amplification succeeds in turning a counterfactual into a fact. No matter how many counterfactuals get entangled, they remain counterfactuals. On the other hand, once a matter of fact exists, it is {\it logically} impossible to erase it. For the relevant matter of fact is not that the needle deflects to the left (in which case one could `erase' it by returning the needle to the neutral position). The relevant matter of fact is that {\it at a time} $t$ the needle deflects (or points) to the left. This is a timeless truth. If at the time $t$ the needle deflects to the left, then it always has been and always will be true that at the time $t$ the needle deflects to the left. Note that an apparatus pointer is not a macroscopic object according to the above definition. In general there is nothing that allows one to predict which way the needle will deflect (given that it will deflect). Only {\it before} and {\it after} the deflection event does the needle behave as a macroscopic object. Is not such a definition self-defeating? It would be so if it were designed to explain why the needle deflects left {\it or} right (rather than both left {\it and} right). But such an explanation is neither required nor possible. If past events allow us to infer a superposition of the form $a\ket{\hbox{left}}\otimes\ket{a}+b\ket{\hbox{right}}\otimes\ket{b}$, they allow us to infer the following: {\it if} there is a matter of fact about the direction in which the needle deflects, it warrants the inference `left' with probability $\absosq{a}$, and it warrants the inference `right' with probability $\absosq{b}$. Nothing allows us to predict the existence of such a matter of fact. The deflection event is a causal primary, notwithstanding that it happens with a measurable probability, and that by a suitable choice of apparatus this probability can be made reasonably large. As I have stressed elsewhere (Mohrhoff, 1999), what is true of particles in double-slit experiments is equally true of cats in double-door experiments. Except for the myriads of matters of fact about the door taken by the cat, `the door taken by the cat' is objectively undefined. This seems to entail a vicious regress. We infer the positions of particles from the positions of the detectors that click. But the positions of detectors are extrinsic, too. They are what they are because of the matters of fact from which one can (in principle) infer what they are. Thus there are detector detectors from which the positions of particle detectors are inferred, and then there are detectors from which the positions of detector detectors are inferred, and so on {\it ad infinitum}. However, as we regress from particle detectors to detector detectors and so on, we sooner or later (sooner rather than later) encounter a macroscopic detector whose position is not manifestly fuzzy. There the buck stops. The positions of things are defined in terms of the not manifestly fuzzy positions of macroscopic objects. It is therefore consistent to think of the deflection of the pointer needle as one of those uncaused actual events on which the (contingent) properties of things depend. {\it Prima facie} we have another vicious regress: Like all contingent properties, the initial and final positions of the needle are what they are because of what happens or is the case in the rest of the world. They thus presupposes other `deflection events,' which presuppose yet other `deflection events,' and so on {\it ad infinitum}. But since before and after its deflection the needle behaves as a macroscopic object, its initial and final positions are quantitatively defined independently of what happens elsewhere. They are positions of the kind that are used to define positions. Hence the deflection event -- the transition from the initial to the final position -- is also independent of what happens elsewhere. \section{\conclu}Language and the Indefinite My chief conclusion in this paper is that (ER1) and (ER2) are both false. The sufficient and necessary condition for the existence of an element of reality $A=a$ is the existence of an actual state of affairs, or the occurrence of an actual event, from which $A=a$ can be inferred. The contingent properties of all quantum systems -- and in clear this means the positions of all material objects and whatever other properties can be inferred from them -- are extrinsic. They are defined in terms of the goings-on in the `rest of the world.' The reason why this does not send us chasing the ultimate property-defining facts in neverending circles, is the existence of a special class of objects the positions of which are not manifestly indefinite. Everything a macroscopic object does (that is, every matter of fact about its present properties) follows via the pertinent classical laws from what it did (that is, from matters of fact about its past properties). This makes it possible to ignore the fact that the properties of a macroscopic object, like all contingent properties, `dangle' from external events and/or states of affairs. Instead of having to conceive of the successive states of a macroscopic object as a bundle of statistically correlated inferences warranted by a multitude of causal primaries external to the object, we are free to think of the object's successive states as an evolving collection of intrinsic properties fastened only to each other, by causal links. The familiar macroworld with its causal links and deterministic processes is something we project onto the fundamental statistical correlations and their uncaused correlata. This projection works where the correlations are not manifestly probabilistic (that is, where the statistical correlations evince no statistical variations). Diachronic correlations that are not manifestly probabilistic can be passed off as causal links. We can impose on them our agent-causality with some measure of consistency, even though this results in the application of a wrong criterion: temporal precedence takes the place of causal independence as the criterion which distinguishes a cause from its effect. Quantum mechanics presupposes the macroworld: it assigns probabilities to conditionals that refer to events or states of affairs either in the actual macroworld or in a possible macroworld. This is the reason why Bohr (1934, 1958) insisted not only on the necessity of describing quantum phenomena in terms of the experimental arrangements in which they are displayed, but also on the necessity of employing classical language in describing these experimental arrangements. Classical language is the language of causal processes, of definite states that evolve deterministically, of definite objects and of definite events -- in short, the language of the macroworld. Thus in one sense the microworld is fundamental (macroscopic objects are made of particles and atoms), and in another sense the macroworld is fundamental (the contingent properties of particles and atoms are defined in terms of the goings-on in the macroworld). The mutual dependence of the quantum and classical `domains' has often been remarked upon (e.g., Landau and Lifshitz, 1977), but I'm not sure it has been adequately appreciated. It seems to me that what is ultimately responsible for this mutual dependence is the conflict between a real, objective indefiniteness and the intrinsic definiteness of language. Language is inherently `classical.' Discourse is of things -- the discrete carriers of significance that appear as the subjects of predicative sentences. Things fall into mutually disjoint classes according to the properties they possess or lack. For any two different classes $C_1$ and $C_2$ there exists a property $p$ such that `$x$ has $p$' is true of all members of $C_1$ and `$x$ lacks $p$' is true of all members of $C_2$. This seems to warrant the following Principle of Completeness (Wolterstorff, 1980): for every thing $x$ and every property $p$, $x$ either has $p$ or lacks $p$. Reality, however, doesn't play along with this linguistic requirement. Sometimes `$x$ has $p$' is neither true nor false but meaningless. There are situations in which nothing in the real world corresponds to the linguistic (or conceptual) distinction between `$x$ has $p$' and `$x$ lacks $p$'. In such situations it is nevertheless meaningful to consider what would have happened if one had found out whether $x$ has $p$ or lacks $p$, and to assign objective probabilities to the alternatives `$x$ has $p$' and `$x$ lacks $p$.' Given the intrinsic definiteness of language, the natural way to express an objective indefiniteness is to use counterfactuals. One then has one counterfactual for each alternative (`if $Q$ were measured, the result would be $q_k$'), and at least one of them comes with a nontrivial objective probability (that is, an objective probability other than 0 or 1). The linguistic requirement of definiteness is met by the use of counterfactuals the respective consequents of which conform to the Principle of Completeness: each consequent explicitly affirms the truth of one alternative and implicitly denies the truth of the other alternatives. The objective indefiniteness finds expression in the fact that the counterfactuals are assigned nontrivial objective probabilities rather than truth values. Objective indefiniteness thus leads to the use of counterfactuals with nontrivial objective probabilities, and nontrivial objective probabilities, as we have seen, entail that the properties affirmed by the counterfactuals' consequents are extrinsic: they are defined in terms of the goings-on in the macroworld, notwithstanding that the objects of the macroworld are made up -- or shall we say, manifested by means -- of nonmacroscopic objects. This mutual dependence of the two `domains' would amount to a vicious circle if the properties of the macroworld were in their turn defined in terms of the microworld. But this is not the case. Since the contingent properties of things are defined in terms of events or states of affairs in the macroworld, quantum mechanics presupposes the macroworld. In particular, it presupposes such matters of fact as `the needle deflects to the left' or `the needle is pointing left.' By itself this does not guarantee that quantum mechanics is consistent with the existence of the macroworld - quantum mechanics (or the interpretation of quantum mechanics put forward in this paper) could lack self-consistency. But self-consistency only requires that the needle's position too is fuzzy, and that it `dangles ontologically' from the goings-on in the rest of the world. Quantum mechanics permits it to `dangle' from them in such a way that, before and after the deflection, it is not manifestly fuzzy. If the needle's position is not manifestly fuzzy, the needle behaves as a macroscopic object, and we can consistently conceive of its successive positions as `dangling causally' from each other -- except for one gap in the causal chain, the deflection event. But being a (probabilistic) transition between states embedded in the causal nexus of the macroworld, this too forms part of the macroworld. Quantum mechanics not only presupposes and admits of the existence of macroscopic objects, it also entails it. The existence of an unpredictable matter of fact about the position of $O$ entails the existence of detectors with `sharper' positions; the existence of an unpredictable matter of fact about the position of one of those detectors entails the existence of detectors with yet `sharper' positions; and so on. It stands to reason that one sooner or later runs out of detectors with `sharper' positions. There are `ultimate' detectors the positions of which are not manifestly fuzzy, and which therefore are macroscopic. \vfill\break \parindent=0pt\bigskip{\bf References}\par\smallskip \baselineskip=14truept\parskip=3truept Aharonov, Y., Bergmann, P.G., and Lebowitz, J.L. (1964) `Time Symmetry in the Quantum Process of Measurement,' {\it Physical Review B} {\bf 134}, 1410-1416. Aharonov,Y., and Vaidman, L. (1991) `Complete Description of a Quantum System at a Given Time,' {\it Journal of Physics} {\bf A 24}, 2315-2328. Albert, D.Z. (1992) {\it Quantum Mechanics and Experience}, Cambridge, MA: Harvard University Press. Bartley III., W.W. (1982) {\it Quantum Theory and the Schism in Physics}, Totowa, NJ: Rowan \& Littlefield. Bell, J.S. (1987) {\it Speakable and Unspeakable in Quantum Mechanics}, Cambridge: Cambridge University Press. Bell, J.S. and Nauenberg, M. (1966) `The Moral Aspect of Quantum Mechanics,' in De Shalit, Feshbach, and Van Hove (1996), pp. 279-86. Reprinted in Bell (1987), pp. 22-28. Bohm, D. (1951) {\it Quantum Theory}, Englewood Cliffs, NJ: Prentice Hall. Bohr, N. (1934) {\it Atomic Theory and the Description of Nature}, Cambridge: Cambridge University Press. Bohr, N. (1958) {\it Atomic Physics and Human Knowledge}, New York: Wiley, p.~72. Bohr, N. (1963) {\it Essays 1958-62 on Atomic Physics and Human Knowledge}, New York: Wiley, p.~3. Cassinello, A., and S\'anchez-G\'omez, J.L. (1996) `On the Probabilistic Postulate of Quantum Mechanics,' {\it Foundations of Physics} {\bf 26}, 1357-1374. Davies, P. (1989) {\it The New Physics}, Cambridge: Cambridge University Press. De Shalit, A., Feshbach, H., and Van Hove, L. (1966) {\it Preludes in Theoretical Physics}, Amsterdam: North Holland. d'Espagnat, B. (1976) {\it Conceptual Foundations of Quantum Mechanics}, 2nd edition, Reading, MA: Benjamin, p. 251. d'Espagnat, B. (1995) {\it Veiled Reality}, Reading, MA: Addison-Wesley. Einstein, A., Podolsky, B., and Rosen, N. (1935) `Can Quantum-Mechanical Description of Physical Reality be Considered Complete?,'' {\it Physical Review} {\bf 47}, 777-780. Englert, B.-G., Scully, M.O., and Walther, H. (1994) `The Duality in Matter and Light,' {\it Scientific American} {\bf 271}, No. 6 (December), 56-61. Englert, B.-G., Scully, M.O., and Walther, H. (1999) `Quantum Erasure in Double-Slit Interferometers with Which-Way Detectors,'' {\it American Journal of Physics} {\bf 67}, 325-329. Ford, J., and Mantica, G. (1992) `Does Quantum Mechanics Obey the Correspondence Principle? Is it Complete?', {\it American Journal of Physics} {\bf 60}, 1068-1097. Gell-Mann, M., and Hartle, J.B. (1990) `Quantum Mechanics in the Light of Quantum Cosmology,' in Zurek (1990), pp. 425-458. Ghirardi, G.C., Rimini, A., and Weber, T. (1986) `Unified Dynamics for Microscopic and Macroscopic Systems,' {\it Physical Review D} {\bf 34}, 470-91. Gisin, N., and Percival, I.C. (1992) `The Quantum-State Diffusion Model Applied to Open Systems,' {\it Journal of Physiccs A} {\bf 25}, 5677-5691. Griffiths, R.B. (1984) `Consistent Histories and the Interpretation of Quantum Mechanics,' {\it Journal of Statistical Physics} {\bf 36}, 219-272. Heisenberg, W. (1958) {\it Physics and Philosophy}, New York: Harper and Row, Chapter 3. Hilgevoord, J. (1998) `The Uncertainty Principle For Energy and Time. II,' {\it American Journal of Physics } {\bf 66}, 396-402. Jauch, J.M. (1968) {\it Foundations of Quantum Mechanics}, Reading, MA: Addison-Wesley. Joos, E., and Zeh, H.D. (1985) `The Emergence of Classical Properties Through Interaction With the Environment,' {\it Zeitschrift f\"ur Physik B} {\bf 59}, 223-243. Kastner, R.E. (1999) `Time-Symmetrized Quantum Theory, Counterfactuals, and ``Advanced Action'',' to be published in {\it Studies in History and Philosophy of Science}. Landau, L.D., and Lifshitz, E.M. (1977) {\it Quantum Mechanics}, Oxford: Pergamon Press, pp.~2-3. Langevin, P. (1939) {Actualit\'es scientifiques et industrielles: Expos\'es de physique g\'en\'erale}, Paris: Hermann. Lockwood, M. (1989) {\it Mind, Brain and the Quantum: The Compound `I'}, Oxford: Basil Blackwell. London, F., and Bauer, E. (1939) `La th\'eorie de l'observation en m\'ecanique quantique,' in Langevin (1939), No. 775. English translation `The Theory of Observation in Quantum Mechanics' in Wheeler and Zurek (1983), pp. 217-259. L\"uders, G. (1951) `\"Uber die Zustands\"anderung durch den Messprozess,' {\it Annalen der Physik (Leipzig)} {\bf 8}, 322-328. Mermin, N.D. (1998) ``What is Quantum Mechanics Trying to Tell Us?,'' {\it American Journal of Physics} {\bf 66}, 753-767. Miller, A.I. (1990) {\it 62 Years of Uncertainty}, New York: Plenum Press. Mohrhoff, U. (1999) `Objectivity, Retrocausation, and the Experiment of Englert, Scully and Walther,' {\it American Journal of Physics} {\bf 67}, 330-335. Mohrhoff, U. (submitted) `What Quantum Mechanics is Trying to Tell Us.' Omn\`es, R. (1992) `Consistent Interpretations of Quantum Mechanics,' {\it Reviews of Modern Physics} {\bf 64}, 339-382. Onley, D., and Kumar, A. (1992) `Time Dependence in Quantum Mechanics -- Study of a Simple Decaying System,' {\it American Journal of Physics} {\bf 60}, 432-439. Page, D.N. (1996) `Sensible Quantum Mechanics: Are Probabilities Only in the Mind?,' {\it International Journal of Modern Physics D} {\bf 5}, 583-596. Page, D.N., and Wootters, W.K. (1983) `Evolution Without Evolution: Dynamics Described by Stationary Observables' {\it Physical Review D} {\bf 27}, 2885-2891. Pais, A., (1982) {\it `Subtle is the Lord...': The Science and the Life of Albert Einstein}, Oxford: Clarendon Press. Pearle, P. (1989) `Combining Stochastic Dynamical State-Vector Reduction with Spontaneous Localization,' {\it Physical Review A} {\bf 39}, 2277-2289. Percival, I.C. (1994) `Primary State Diffusion,' {\it Proceedings of the Royal Society London} {\bf A447}, 189-209. Peres, A., and Zurek, W.H. (1982) `Is Quantum Theory Universally Valid?,' {\it American Journal of Physics} {\bf 50}, 807-810. Popper, K.R. (1982) in Bartley III (1982). Price, H. (1996) {\it Time's Arrow \& Archimedes' Point}, New York: Oxford University Press. Primas, H. (1990) `Mathematical and Philosophical Questions in the Theory of Open and Macroscopic Quantum Systems,' in Miller (1990), pp. 233-257. Redhead, M. (1987) {\it Incompleteness, Nonlocality and Realism}, Oxford: Clarendon, p. 72. Shimony, A. (1978) `Metaphysical Problems in the Foundations of Quantum Mechanics,' {\it International Philosophical Quarterly} {\bf 18}, 3-17. Shimony, A. (1989) `Conceptual Foundations of Quantum Mechanics' in Davies (1989), 373-95. Scully, M.O., Englert, B.-G., and Walther, H. (1991) `Quantum Optical Tests of Complementarity,' {\it Nature} {\bf 351}, No. 6322, 111-116. Singh, I., and Whitaker, M.A.B. (1982) `Role of the Observer in Quantum Mechanics and the Zeno Paradox,' {\it American Journal of Physics} {\bf 50}, 882-887. Stapp, H.P. (1993) {\it Mind, Matter, and Quantum Mechanics}, Berlin: Springer. Vaidman, L. (1993) `Lorentz-Invariant ``Elements of Reality'' and the Joint Measurability of Commuting Observables,' {\it Physical Review Letters} {\bf 70}, 3369-3372. Vaidman, L. (1996a) `Defending Time-Symmetrized Quantum Theory,' e-print archive quant-ph 9609007. Vaidman, L. (1996b) `Weak-Measurement Elements of Reality,' {\it Foundations of Physics} {\bf 26}, 895-906. Vaidman, L. (1997) `Time-Symmetrized Counterfactuals in Quantum Theory,' e-print archive quant-ph 9807075, Tel-Aviv University Preprint TAUP-2459-97. Vaidman, L. (1998) `Time-Symmetrized Quantum Theory,' {\it Fortschritte der Physik} {\bf 46}, 729-739. Vaidman, L. (1999) `Defending Time-Symmetrized Quantum Counterfactuals,' to be published in {\it Studies in History and Philosophy of Science}. von Neumann, J. (1955) {\it Mathematical Foundations of Quantum Mechanics}, Princeton: Princeton University Press. Wheeler, J.A. (1983) `On Recognizing ``Law Without Law'' (Oersted Medal Response at the Joint APS-AAPT Meeting, New York, 25 January 1983),' {\it American Journal of Physics} {\bf 51}, 398-404. Wheeler, J.A., and Zurek, W.H. (1983) {\it Quantum Theory and Measurement}, Princeton, NJ: Princeton University Press. Wolterstorff, N. (1980), {\it Works and Worlds of Art}, Oxford: Clarendon Press. Zurek, W.H., (1981) `Pointer Basis of Quantum Apparatus: Into What Mixture Does the Wave Packet Collapse?', {\it Physical Review D} {\bf 24}, 1516-1525. Zurek, W.H., (1982) `Environment-Induced Superselection Rules,' {\it Physical Review D} {\bf 26}, 1862-1880. Zurek, W.H., (1990) {\it Complexity, Entropy, and the Physics of Information}, Reading, MA: Addison-Wesley. \bye
\section*{Acknowledgements} St. W. acknowledges financial support by the {\em Schweizerische Nationalfonds}.
\section{Introduction} Nature appears to be covariant both under the discrete transformation CPT, formed from the product of charge conjugation C, parity inversion P, and time reversal T, and under the continuous Lorentz transformations including rotations and boosts. The CPT theorem links these symmetries, stating that under mild technical assumptions CPT is an exact symmetry of local Lorentz-covariant field theories of point particles.\cite{cpt,sachs} High-precision tests of both CPT and Lorentz invariance exist. According to the Particle Data Group\cite{pdg} the best figure of merit for CPT tests involves the kaon particle-antiparticle mass difference, which has been bounded by experiments at Fermilab and CERN to\cite{kexpt} \begin{equation} \fr{|m_K - m_{\overline K}|}{m_K} \mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-18} \quad . \label{a} \end{equation} Indeed, at present CPT is the only combination of C, P, T observed as an exact symmetry of nature at the fundamental level. The existence of high-precision experimental tests and of the general CPT theorem for Lorentz-covariant particle theories means that the observation of CPT or Lorentz violation would be a sensitive signal for unconventional physics beyond the standard model. It is therefore interesting to consider possible theoretical mechanisms through which CPT or Lorentz symmetry might be violated. Most suggestions along these lines in the literature either have physical features that seem unlikely to be realized in nature or involve radical revisions of conventional quantum field theory, or both. Nonetheless, there does exists at least one promising theoretical possibility, based on spontaneous breaking of CPT and Lorentz symmetry in an underlying theory,\cite{ks,kp} that appears to be compatible both with experimental constraints and with established quantum field theory. It suggests that apparent breaking of CPT and Lorentz symmetry might be observable in existing or feasible experiments, and it leads to a general phenomenology for CPT and Lorentz violation at the level of the standard model and quantum electrodynamics (QED). The formulation and experimental implications of this theory are briefly described in this talk. \section{Framework} In principle, one can attempt to circumvent the difficult issue of developing a satisfactory microscopic theory allowing CPT and Lorentz breaking by adopting a purely phenomenological approach. This can be done by identifying and parametrizing observable quantities that allow for CPT or Lorentz violation. A well-known example is the phenomenology of CPT violation in oscillations of neutral kaons.\cite{leewu} In the neutral-kaon system, linear combinations of the strong-interaction eigenstates $K^0$ and $\overline{K^0}$ form the physical eigenstates $K_S$ and $K_L$. These combinations contain two complex parameters, $\epsilon_K$ and $\delta_K$, parametrizing CP violation. One, $\epsilon_K$, governs T violation with CPT symmetry while the other, $\delta_K$, governs CPT violation with T symmetry. The standard model of particle physics has a mechanism for T violation, and so $\epsilon_K$ is in this context nonzero and in principle calculable. However, CPT is a symmetry of the standard model and so $\delta_K$ is expected to vanish. The possibility of a nonzero value of $\delta_K$ is from this prespective only a phenomenological choice. It has no grounds in a microscopic theory and $\delta_K$ is therefore not calculable. Indeed, in the absence of a microscopic theory, it is even unclear whether this parametrization makes physical sense. Moreover, without a microscopic origin $\delta_K$ cannot be linked to other phenomenological parameters for CPT tests in different experiments. Evidently, it is more attractive theoretically to develop an explicit microscopic theory for CPT and Lorentz violation. With a theory of sufficient scope, a general and quantitative phenomenology for CPT and Lorentz violation could then be extracted at the level of the standard model. This would allow calculation of phenomenological parameters, direct comparisons between experiments, and perhaps the prediction of signals. The development of a microscopic theory of this type is feasible within the context of spontaneous CPT and Lorentz breaking.\cite{ks,kp} The idea is that the underlying theory of nature has a Lorentz- and CPT-covariant action, but apparent violations of these symmetries could result from their spontanteous violation in solutions to the theory. It appears that this mechanism is viable from the theoretical viewpoint and is an attractive way to violate CPT and Lorentz invariance. Since spontaneous breaking is a property of the solution rather than the dynamics of a theory, the broken symmetry plays an important role in establishing the physics. In the case of CPT and Lorentz violation, spontaneous breaking has the advantage that many of the desirable properties of a Lorentz-covariant theory can be expected. This is in sharp distinction to other types of CPT and Lorentz breaking, which often are inconsistent with theoretical notions such as causality or probability conservation. The physics of a particle in a vacuum with spontaneous Lorentz violation is in some respects similar to that of a conventional particle moving inside a biaxial crystal.\cite{cksm} This system typically breaks Lorentz covariance both under rotations and under boosts. However, instead of leading to fundamental problems, the lack of Lorentz covariance is merely a result of the presence of the background crystal fields, which leaves unaffected features such as causality. Indeed, one can explicitly confirm microcausality in certain simple models arising from spontaneous CPT and Lorentz breaking.\cite{cksm} In a Lorentz-covariant theory, certain types of interaction among Lorentz-tensor fields could trigger spontaneous breaking of Lorentz symmetry. The idea is that these interactions could destabilize the naive vacuum and generate nonzero Lorentz-tensor expectation values, which fill the true vacuum and cause spontaneous Lorentz breaking.\cite{ks} This also induces spontaneous CPT violation whenever the expectation values involve tensor fields with an odd number of spacetime indices.\cite{kp} Provided components of the expectation values lie along the four macroscopic spacetime dimensions, apparent violations of CPT and Lorentz symmetry could arise at the level of the standard model.\cite{kp2} This could lead to observable effects, some of which are described in the following sections. Conventional four-dimensional renormalizable gauge theories such as the standard model lack the necessary destabilizing interactions to trigger spontaneous Lorentz violation. However, the mechanism may be realized in some string (M) theories because suitable Lorentz-tensor interactions occur. This can be investigated using string field theory in the special case of the open bosonic string, where the action and equations of motion can be analytically derived for particle fields below some fixed level number $N$. Obtaining and comparing solutions for different $N$ allows the identification of solutions that persist as $N$ increases.\cite{ks} For some cases this procedure has been performed to a depth of over 20,000 terms in the static potential.\cite{kp} The solutions remaining stable as $N$ increases include ones spontaneously breaking Lorentz symmetry. In standard field theories, spontaneous breaking of a continuous global symmetry is accompanied by the appearance of massless modes, ensured by the Nambu-Goldstone theorem. Promoting a global spontaneously broken symmetry to a local gauge symmetry leads to the Higgs mechanism: the massless modes disappear and a mass is generated for the gauge boson. Similarly, spontaneous breaking of a continuous global Lorentz symmetry would also lead to massless modes. However, although the inclusion of gravity promotes Lorentz invariance to a local symmetry, no analogue to the Higgs effect occurs.\cite{ks} The dependence of the connection on derivatives of the metric rather than the metric itself ensures that the graviton propagator is affected in such a way that no graviton mass is generated when local Lorentz symmetry is spontaneously broken. \section{Standard-Model and QED Extensions} Assuming spontaneous CPT and Lorentz violation does occur, then any apparent breaking at the level of the SU(3)$\times$SU(2)$\times$U(1) standard model and QED must be highly suppressed to remain compatible with established experimental bounds. If the appropriate dimensionless suppression factor is determined by the ratio of a low-energy (standard-model) scale to the (Planck) scale of an underlying fundamental theory, then relatively few observable effects of Lorentz or CPT violation would arise. To study these, it is useful to develop an extension of the standard model obtained as the low-energy limit of the fundamental theory.\cite{kp2} To gain insight about the construction of such an extension, consider as an example a possible coupling between one or more bosonic tensor fields and fermion bilinears in the low-energy limit of the underlying theory. When the tensors acquire expectation values $\vev{T}$, the low-energy theory gains additional terms of the form \begin{equation} {\cal L} \sim \fr {\lambda} {M^k} \vev{T}\cdot\overline{\psi}\Gamma(i\partial )^k\chi + {\textstyle h.c.} \quad . \label{aa} \end{equation} Here, the gamma-matrix structure $\Gamma$ and the $k$ spacetime derivatives $i\partial$ determine the Lorentz properties of the bilinear in the fermion fields $\psi$, $\chi$ and hence fix the type of apparent CPT and Lorentz violation in the low-energy theory. The effective coupling involves an expectation $\vev{T}$ together with a dimensionless coupling $\lambda$ and a suitable power of a large (Planck or compactification) scale $M$ associated with the fundamental theory. Proceeding along these lines, one can determine all possible terms arising at the level of the standard model from spontaneous CPT and Lorentz breaking in any underlying theory (not necessarily string theory). This leads to a general Lorentz-violating extension of the standard model that includes both CPT-even and CPT-odd terms.\cite{cksm} It contains all possible allowed hermitian terms preserving both SU(3)$\times$SU(2)$\times$U(1) gauge invariance and power-counting renormalizability. It appears at present to be the sole candidate for a consistent extension of the standard model based on a microscopic theory of Lorentz violation. Despite the apparent CPT and Lorentz breaking, the standard-model extension exhibits several desirable properties of conventional Lorentz-covariant field theories by virtue of its origin in spontaneous symmetry breaking from a covariant underlying theory.\cite{cksm} Thus, the usual quantization methods are valid and features like microcausality and positivity of the energy are to be expected. Also, energy and momentum are conserved provided the tensor expectation values are independent of spacetime position (no soliton solutions). Even one type of Lorentz symmetry remains: the theory is covariant under rotations or boosts of the observer's inertial frame (observer Lorentz transformations). The apparent Lorentz violations appear only when (localized) fields are rotated or boosted (particle Lorentz transformations) relative to the vacuum tensor expectation values. In the case of the conventional standard model, one can obtain the usual versions of QED by taking suitable limits. For the standard-model extension, it can be shown that the usual gauge symmetry breaking to the electromagnetic U(1) occurs, and taking appropriate limits yields generalizations of the usual versions of QED. It turns out that the apparent CPT and Lorentz breaking can arise in both the photon and fermion sectors.\cite{cksm} These extensions of QED are of particular interest because many high-precision QED tests of CPT and Lorentz symmetry exist. An explicit and relatively simple example is the restriction of the standard-model extension to an extension of QED involving only photons, electrons, and positrons. The usual lagrangian is: \begin{equation} {\cal L}^{\rm QED} = \overline{\psi} \gamma^\mu ({\textstyle{1\over 2}} i \stackrel{\leftrightarrow}{\partial_\mu} - q A_\mu ) \psi - m \overline{\psi} \psi - \frac 1 4 F_{\mu\nu}F^{\mu\nu} \quad . \label{aaa} \end{equation} Apparent Lorentz violation can occur in both the fermion and photon sectors, and it can be CPT even or CPT odd. The CPT-violating terms are: \begin{equation} {\cal L}^{\rm CPT}_{e} = - a_{\mu} \overline{\psi} \gamma^{\mu} \psi - b_{\mu} \overline{\psi} \gamma_5 \gamma^{\mu} \psi \quad , $$ $$ {\cal L}^{\rm CPT}_{\gamma} = {\textstyle{1\over 2}} (k_{AF})^\kappa \epsilon_{\kappa\lambda\mu\nu} A^\lambda F^{\mu\nu} \quad . \label{bbb} \end{equation} The CPT-preserving terms are: \begin{equation} {\cal L}^{\rm Lorentz}_{e} = c_{\mu\nu} \overline{\psi} \gamma^{\mu} ({\textstyle{1\over 2}} i \stackrel{\leftrightarrow}{\partial^\nu} - q A^\nu ) \psi + d_{\mu\nu} \overline{\psi} \gamma_5 \gamma^\mu ({\textstyle{1\over 2}} i \stackrel{\leftrightarrow}{\partial^\nu} - q A^\nu ) \psi - {\textstyle{1\over 2}} H_{\mu\nu} \overline{\psi} \sigma^{\mu\nu} \psi $$ $$ {\cal L}^{\rm Lorentz}_{\gamma} = -\frac 1 4 (k_F)_{\kappa\lambda\mu\nu} F^{\kappa\lambda}F^{\mu\nu} \quad . \label{ccc} \end{equation} The reader is referred to the literature\cite{cksm} for details of the notation and conventions and for information about the properties of the extra terms. Note, however, that all these terms are invariant under observer Lorentz transformations, whereas the expressions in Eqs.\ (4) and (5) violate particle Lorentz invariance: the coefficients of the extra terms behave as (minuscule) Lorentz- and CPT-violating couplings. Note also that not all the components of the coefficients appearing are physically observable. For example, field redefinitions can be used to eliminate some coefficients of the type $a_\mu$ in the standard-model extension. It turns out that these can be directly detected only in flavor-changing experiments, and so they are unobservable at leading order in experiments restricted to electrons, positrons, and photons. \section{Experimental Tests} The standard-model extension described above forms a quantitative framework within which various experimental tests of CPT and Lorentz symmetry can be studied and compared. Moreover, potentially observable signals can be deduced in some cases. Evidently, any tests seeking to establish nonzero CPT- and Lorentz-violating terms in the standard-model extension must contend with the expected heavy suppression of physical effects. Although many tests of CPT and Lorentz symmetry lack the necessary sensitivity to possible signals, a few special ones can already place useful constraints on some of the new couplings in the standard-model extension. Several of these tests are discussed elsewhere in these proceedings. Among the ones investigated to date are experiments with neutral-meson oscillations,\cite{kexpt,kp,kp2}$^{\!-\,}$\cite{ak} comparative tests of QED in Penning traps,\cite{pennexpts,bkr} spectroscopy of hydrogen and antihydrogen,\cite{antih,bkr2} measurements of cosmological birefringence,\cite{cksm} and observations of the baryon asymmetry.\cite{bckp} The remainder of this talk provides a brief outline of some of these studies. Other work is in progress, including an investigation\cite{kla} of constraints from clock-comparison experiments.\cite{cc} \subsection{Neutral-Meson Oscillations} Flavor oscillations occur or are anticipated in a variety of neutral-meson systems, including $K$, $D$, $B_d$, and $B_s$. A neutral-meson state evolves in time according to a non-hermitian two-by-two effective hamiltonian in the meson-antimeson state space. The effective hamiltonian involves complex parameters $\epsilon_P$ and $\delta_P$ that govern (indirect) CP violation, where the neutral meson is denoted by $P$. In the $K$ system, $\epsilon_K$ and $\delta_K$ are the same phenomenological quantities mentioned in section 2. The parameter $\epsilon_P$ governs T violation, while $\delta_P$ governs CPT violation. Bounds on CPT violation can be obtained by constraining the magnitude of $\delta_P$ in experiments with meson oscillations. In the context of the usual standard model, which preserves CPT, $\delta_P$ is necessarily zero. In contrast, in the context of the standard-model extension $\delta_P$ is a derivable quantity.\cite{ak} It turns out that at leading order $\delta_P$ depends only on a single type of extra coupling in the standard-model extension. This type of coupling has the form $- a^q_{\mu} \overline{q} \gamma^\mu q$, where $q$ represents one of the valence quark fields in the $P$ meson and the quantity $a^q_{\mu}$ is spacetime constant but depends on the quark flavor $q$. Since Lorentz symmetry is broken in the standard-model extension, the derived expression for $\delta_P$ varies with the boost and orientation of the $P$ meson.\cite{ak} Denoting by $\beta^\mu \equiv \gamma(1,\vec\beta)$ the four-velocity of the $P$-meson in the frame in which the quantities $a^q_{\mu}$ are specified, it can be shown that $\delta_P$ is given at leading order in all coupling coefficients in the standard-model extension by \begin{equation} \delta_P \approx i \sin\hat\phi \exp(i\hat\phi) \gamma(\Delta a_0 - \vec \beta \cdot \Delta \vec a) /\Delta m \quad . \label{e} \end{equation} For simplicity, subscripts $P$ have been omitted on the right-hand side. In Eq.\ \rf{e}, $\Delta a_\mu \equiv a_\mu^{q_2} - a_\mu^{q_1}$, where $q_1$ and $q_2$ are the valence-quark flavors for the $P$ meson. Also, $\hat\phi\equiv \tan^{-1}(2\Delta m/\Delta\gamma)$, where $\Delta m$ and $\Delta \gamma$ are the mass and decay-rate differences, respectively, between the $P$-meson eigenstates. This result has several implications. One is that tests of CPT and Lorentz symmetry with neutral mesons are independent at leading order of all other types of tests mentioned here. This is because $\delta_P$ is sensitive only to $a^q_{\mu}$ and because this sensitivity arises from flavor-changing effects. None of the other experiments described here involve flavor changes, which can be shown to imply that none are sensitive to any $a^q_{\mu}$. The result \rf{e} also makes predictions about signals in experiments with neutral mesons. For example, the real and imaginary parts of $\delta_P$ are predicted to be proportional.\cite{kp2} Similarly, Eq.\ \rf{e} suggests that the magnitude of $\delta_P$ may be different for different $P$ due to the flavor dependence of the coefficients $a_\mu^q$. For example, if the coefficients $a_\mu^q$ grow with mass as do the usual Yukawa couplings, then the heavier neutral mesons such as $D$ or $B_d$ may exhibit the largest CPT-violating effects. The dependence of the result \rf{e} on the meson boost magnitude and orientation implies several notable effects in the signals for CPT and Lorentz violation.\cite{ak} For example, two different experiments may have inequivalent CPT and Lorentz reach despite having comparable statistical sensitivity. This could arise if the mesons for one experiment are well collimated while those for the other have a $4\pi$ distribution, or if the mesons involved in the two experiments have very different momentum spectra. Another interesting effect is the possibility of diurnal variations in the data, arising from the rotation of the Earth relative to the orientation of the coupling coefficients.\cite{ak} This issue may be of some importance because the data in neutral-meson experiments are typically taken over many days. At present, the tightest clean experimental constraints on CPT violation come from observations of the $K$ system.\cite{kexpt} Some experimental results are now also available for the heavier neutral-meson systems. Two collaborations\cite{bexpt} at CERN have performed analyses to investigate whether\cite{ckv} existing data suffice to bound CPT violation. The OPAL collaboration has published the measurement $\hbox{Im}\,\delta_{B_d} = -0.020 \pm 0.016 \pm 0.006$, while the DELPHI collaboration has announced a preliminary result of $\hbox{Im}\,\delta_{B_d} = -0.011 \pm 0.017 \pm 0.005$. Further theoretical and experimental studies are underway. \subsection{QED Experiments} High-precision measurements of properties of particles and antiparticles can be obtained by trapping individual particles for extended time periods. Comparison of the results provides sensitive CPT tests. Such experiments can constrain the couplings in the fermion sector of the QED extension.\cite{bkr} Penning traps can be used to obtain comparative measurements of particle and antiparticle anomaly and cyclotron frequencies.\cite{pennexpts} The QED extension predicts direct signals and also effects arising from diurnal variations in the Earth-comoving laboratory frame.\cite{bkr} Appropriate figures of merit for the various signals have been defined and the attainable experimental sensitivity estimated. As one example, comparing the anomalous magnetic moments of electrons and positrons would generate an interesting bound on the spatial components of the coefficient $b^e_\mu$ in the laboratory frame. Available technology could place a limit of order $10^{-20}$ on the associated figure of merit. A related test involves the search for diurnal variations of the electron anomaly frequency, for which a new experimental result with a figure of merit bounded at $6\times 10^{-21}$ is presented in these proceedings by Mittleman, Ioannou, and Dehmelt.\cite{rm} Analogous experiments with protons and antiprotons may be feasible. Particle and antiparticle cyclotron frequencies can also be compared. In these proceedings, Gabrielse and coworkers present the results of an experiment comparing the cyclotron frequencies of $H^-$ ions and antiprotons in the same trap.\cite{gk} The leading-order effects in this experiment provide a test of Lorentz violation in the context of the standard-model extension, with an associated figure of merit bounded at $4\times 10^{-25}$. Tests of CPT and Lorentz symmetry are also possible via high-precision comparisons of spectroscopic data from trapped hydrogen and antihydrogen.\cite{antih} An investigation of the possible experimental signals within the context of the standard-model and QED extensions has been performed.\cite{bkr2} Direct sensitivity to CPT- and Lorentz-violating couplings, without suppression factors associated with the fine-structure constant, arises for certain specific 1S-2S and hyperfine transitions in magnetically trapped hydrogen and antihydrogen. In principle, theoretically clean signals might be observed for particular types of CPT and Lorentz violation. The photon sector of the QED extension can also be tightly constrained from a combination of theoretical considerations and terrestrial, astrophysical, and cosmological experiments on electromagnetic radiation. It is known that the pure-photon CPT-violating term in Eq.\ \rf{bbb} can generate negative contributions to the energy,\cite{cfj} which may limit its viability and suggests the coefficient $(k_{AF})^\kappa$ should be zero.\cite{cksm} In contrast, the CPT-even term in the following equation maintains a positive conserved energy.\cite{cksm} The solutions of the extended Maxwell equations with CPT- and Lorentz-breaking effects involve two independent propagating degrees of freedom,\cite{cksm} as usual. Unlike the conventional propagation of electromagnetic waves in vacuum, however, in the extended Maxwell case the two modes have different dispersion relations. This means the vacuum is birefringent. Indeed, the effects of the CPT and Lorentz violation on an electromagnetic wave traveling in the vacuum are closely analogous to those exhibited by an electromagnetic wave in conventional electrodynamics that is passing through a transparent optically anisotropic and gyrotropic crystal with spatial dispersion of the axes.\cite{cksm} The sharpest experimental limits on the extra coefficients in the extended Maxwell equations can be obtained by constraining the birefringence of radio waves on cosmological distance scales. Considering first the CPT-odd coefficient $(k_{AF})_\mu$, one finds\cite{cfj,hpk} a bound of the order of $\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-42}$ GeV on its components. A disputed claim exists\cite{nr,misc} for a nonzero effect at the level of $|\vec k_{AF}|\sim 10^{-41}$ GeV. For the CPT-even dimensionless coefficient $(k_F)_{\kappa\lambda\mu\nu}$, the single rotation-invariant irreducible component is constrained to $\mathrel{\rlap{\lower4pt\hbox{\hskip1pt$\sim$} 10^{-23}$ by the existence of cosmic rays\cite{cg} and other tests. Rotation invariance is broken by all the other irreducible components of $(k_F)_{\kappa\lambda\mu\nu}$. Although in principle it might be feasible to constrain these coefficients with existing techniques for measuring cosmological birefringence,\cite{cksm} no limits presently exist. It is plausible that a bound at the level of about $10^{-27}$ could be placed on components of $(k_F)_{\kappa\lambda\mu\nu}$. The sharp experimental constraints obtained on $(k_{AF})_\mu$ are compatible with the zero value needed to avoid negative-energy contributions. However, no symmetry protects a zero tree-level value of $(k_{AF})_\mu$. It might therefore seem reasonable to expect $(k_{AF})_\mu$ to acquire a nonzero value from radiative corrections involving CPT-violating couplings in the fermion sector. Nonetheless, this does \it not \rm occur:\cite{cksm} an anomaly-cancellation mechanism can ensure that the net sum of all one-loop radiative corrections is finite. The situation is technically involved because the contribution from each individual radiative correction is ambiguous,\cite{cksm,jk} but the anomaly-cancellation mechanism can hold even if one chooses to define the theory such that each individual radiative correction is nonzero. Thus, a tree-level CPT-odd term is unnecessary for one-loop renormalizability. Similar effects may occur at higher loops. This ability to impose the vanishing of an otherwise allowed CPT-odd term represents a significant check on the consistency of the standard-model extension. For the CPT-even Lorentz-violating pure-photon term there is no similar mechanism, and in fact calculations have explicitly demonstrated\cite{cksm} the existence of divergent radiative corrections at the one-loop level. This therefore leaves open the interesting possibility of future detection of a nonzero effect via measurements of cosmological birefringence. Various other possible observable CPT effects have been identified. For example, under suitable conditions the observed baryon asymmetry can be generated in thermal equilibrium through CPT- and Lorentz-violating bilinear terms.\cite{bckp} A relatively large baryon asymmetry produced at grand-unified scales would eventually become diluted to the observed value through sphaleron or other effects. This mechanism represents one possible alternative to the conventional scenarios for baryogenesis, in which nonequilibrium processes and C- and CP-breaking interactions are required.\cite{ads} \section*{Acknowledgments} I thank Orfeu Bertolami, Robert Bluhm, Chuck Lane, Don Colladay, Roman Jackiw, Rob Potting, Neil Russell, Stuart Samuel, and Rick Van Kooten for collaborations. This work was supported in part by the United States Department of Energy under grant number DE-FG02-91ER40661. \section*{References}
\section{Introduction} \label{sec:intro} Neutral gas at high redshifts is most easily observed through the Lyman-$\alpha$ transition of the hydrogen atom, which with current technology can be detected in absorption against the UV continuum of QSOs even at column densities as low as low as $10^{13}$ atoms/cm$^{-2}$. The bulk (by mass) of the neutral gas however is found in the few very high column density systems(Rao and Briggs, 1993), where one could in principle expect a non trivial molecular fraction. However, quantitative predictions of the molecular fraction are difficult to make since the conversion of gas from atomic to molecular form depends on a variety of environmental factors like the UV background, the metallicity and the dust content, all of which are poorly constrained at high redshift. On the observational front, despite searching a large sample, mm molecular lines have been detected in absorption at high redshifts only from four sources (of which two are gravitational lenses and two appear to arise from gas associated with the AGN itself) (Wiklind \& Combes 1998). Here we discuss the case of PKS~1830-21, which is the brightest known radio lens. PKS~1830-21 was identified as a candidate gravitational lens on the basis of its peculiar radio spectrum and morphology (Rao \& Subhramanyan 1988, Subhramanyan et. al. 1990, Jauncey et. al. 1991). The radio structure (see Figure~\ref{fig:ov}) consists of two compact flat spectrum components separated by $\sim 1^{''}$ (henceforth called the northeast (NE) and southwest (SW) components respectively), joined by a steep spectrum ring. At a frequency of 1.7~GHz roughly one third of the observed flux comes from the ring and each of two compact components. At the redshifted frequencies of HI (753 MHz) and OH (884 MHz) the ring is expected to be even more dominant. The lack of simultaneous multi-frequency flux density measurements of sufficient angular resolution (in view of the strong variability of 1830-21) makes a more accurate assessment of the ring flux and the relative components fluxes at the low frequencies not possible at present. For long, no optical counter part has been found for 1830-21 (Djorgovski et. al. 1992), largely because of confusion arising from its low galactic latitude, although there is now some evidence for one (Courbin et. al. 1998). Two independent gravitational lensing models have been proposed for PKS~1830-21, (Nair et. al. 1993, Kochanek \& Narayan 1992). At the time that these models were made no redshift was available either for the source or the lens. The redshift of the lens is now known to be $0.89$ from molecular line observations (Wiklind \& Combes 1996). The absorption spectra against the NE and the SW image are very different (Frye et. al. 1996, Wiklind \& Combes 1998), ruling out the possibility that the molecules at $0.89$ are associated with the background quasar itself. The bulk of the molecular absorption occurs against the SW component, although much weaker absorption is also seen in some molecules against the NE component. The velocity separation between the absorption seen against the NE image and the SW image is 147 km/s. In addition to the molecules seen at $z=0.88582$, HI absorption has also been seen towards PKS~1830-21, but at a lower reshift of $0.19$ (Lovell et. al. 1996). The velocity width of this HI line is $\sim 30$ km/s and it has been interpreted as arising due to absorption in a dense spiral arm of a low redshift spiral galaxy. No molecular absorption has been detected from this lower redshift system (Wiklind \& Combes 1997). In what follows we report on WSRT observations of the HI and OH absorption arising from the system at $z=0.89$. At mm wavelenghts only the extremely compact, flat spectrum components of the background source have sizeable flux. Consequently the spectra sample a region of order only a few tens of parsecs across. At the HI and OH frequencies however, the background source is considerably more extended. These lines are thus more suited to probe the large scale kinematics of the absorbing system as well as to determine the averaged physical properties on a kpc scale. \section{ Observations and data reduction } \label{sec:obs} The observations were done with the broad band UHF receivers installed at the WSRT as part of the on going WSRT upgrade. The HI observations are summarized in Table~\ref{tab:hi} and the OH observations in Table~\ref{tab:oh} \begin{table} \caption[]{HI Observations} \label{tab:hi} \begin{tabular}{lcc} \hline\hline Date & Bandwidth & Channel Separation \\ & MHz (km/s) & (km/s) \\ \hline 03/Nov/96 & 2.5 (996) & 31 \\ 15/Nov/96 & 5.0 (1992) & 31 \\ 08/Jan/97 & 2.5 (996) & 1 \\ \hline \end{tabular} \end{table} \begin{table} \caption[]{OH Observations} \label{tab:oh} \begin{tabular}{lcc} \hline\hline Date & Bandwidth & Channel Separation \\ & MHz (km/s)$^{\dagger}$ & (km/s)$^{\dagger}$ \\ \hline 17/Nov/96 & 5.0 (1696) & 26.5 \\ 15/Dec/96 & 5.0 (1696) & 26.5 \\ \hline \end{tabular} \begin{list}{}{} \item[$^{\dagger}$] The velocity scale is for the 1667~MHz transition. \end{list} \end{table} The OH observations were made using the standard interferometric mode and the data were reduced using NEWSTAR, the WSRT data reduction package. 1830-21 is spatially unresolved at the WSRT baselines. The data from the two observing runs were added together (after applying the appropriate heliocentric Doppler correction) and is shown in Figure~\ref{fig:oh}. In addition, a lower resolution but larger total bandwidth spectrum was also obtained. This spectrum (which is not included here) is substantially the same as that shown in Figure 2. No broader absorption features were detected. The high resolution HI spectrum, Figure 1{c} was obtained using the WSRT as a compound interferometer~(CI), where the telescope was divided into two phased arrays and the output of these phased arrays was fed into the correlator. This mode achieves high spectral resolution at the expense of losing spatial information. However since PKS~1830-21 is not resolved at the WSRT, there is no loss of spatial information in the CI mode. The CI data was reduced using the WASP package (Chengalur 1996). The spectrum agrees well with that of Carilli et. al. (1997), apart from the region near $v\sim 0$, where their spectrum is badly affected by interference. The line is fully resolved and reaches a peak optical depth of 5.5\%. The lower resolution HI spectra Figure 1a\&b were obtained in the standard interferometric mode and reduced using NEWSTAR. The observation on 15/Nov/96 used a much larger bandwidth, however again no new broad absorption feature was detected. As in the case for OH (but this time with better sensitivty and a longer time baseline), there is no measureable difference between the spectra obtained over a period of $\sim 2$ months. The flux densities were calibrated via reference to 3C48 for which we adopt a flux of 25.5~Jy at 753~MHz and 22.7~Jy at 884~MHz, which are based on the Baars et al. (1997) scale. \begin{figure} \resizebox{\hsize}{!}{\includegraphics{hi.eps}} \caption[]{HI spectrum at $z=0.89$ towards PKS~1830-21. {\bf a}~The velocity resolution is $31$~km/s. The velocity scale here and throughout the paper is with respect to the molecular absorption towards the SW image at $z_{hel}=0.88582$. The solid line is a two gaussian fit to the spectrum, the component gaussians are shown by dotted and dashed lines respectively. The position and FWHM of the narrow gaussian are $-148$~km/s and $40$~km/s respectively. The position and FWHM of the broader gaussian is $-80$~km/s and $120$~km/s respectively. {\bf b}~The HI spectrum again at a resolution of $31$~km/s, but with a total bandwidth of $\sim 2000$~km/s. No broad absorption is seen, nor are any narrow absorption components seen at large velocites. The overlaid solid line is the fit to spectrum in part {\bf a}. {\bf c}~The CI spectrum with a resolution of $\sim 1$~km/s. Once again the superimposed solid line is not a separate fit, but the same fit as in panel {\bf a}. No new narrow absorption features are seen. The continuum flux of 1830-211 at the observed frequency is about 10.5 Jy} \label{fig:hi} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{oh.eps}} \caption[]{OH spectrum at $z=0.89$ towards PKS~1830-21. The lower axis shows the velocity scale (with respect to $z=0.88582$) for the 1667~MHz transition, while the velocity scale for the 1665~MHz transition is shown in the upper axis. The velocity resolution is $\sim 27$~km/s. The continuum flux at the observed frequency of 884 Mhz is about 10.1 Jy} \label{fig:oh} \end{figure} \begin{figure} \resizebox{\hsize}{!}{\includegraphics{ov1.ps}} \caption[]{Schematic showing PKS~1830-21 superposed on a typical galactic velocity field. The galaxy inclination and position angle are close to that of the best fit model in Nair et. al. 1993. The radio contours are from the 6cm Merlin image of Patnaik et al. 1994. } \label{fig:ov} \end{figure} \section{ Discussion } \label{sec:dis} With peak optical depths of only 0.007 and 0.005 in the two OH absorption lines the profile shape is not so well defined as that of the HI line. However, it is clear that the OH spectrum and the HI spectrum have similar overall velocity widths. Since the separation of the two OH lines is 350 km/sec we conclude that they do not overlap, consistent with the height of the continuum inbetween the two absorption features. The 1667~MHz transition has an integrated optical depth that is larger than that of the 1665~MHz transition. Within the measurement errors the ratio of the optical depth is consistent with the 9:5 ratio expected in thermal equilibrium. There is evidence that at zero velocity the 1665 MHz line is deeper than the 1667 MHz line, suggesting variations in opacity of the 1665/1667 transitions. This could be related to the much larger molecular line optical depth at zero velocity than at -147 km/sec. We hope to address this issue with future more sensitive observations of the OH lines. The optical depth ($\sim 10^{-2}$), the velocity width, the overall optical depth ratio, and the ratio of the OH column density to the excitation temperature ($N_{OH}/T_{ex} \sim 4 \times 10^{14}$) are all within the range of OH absorption seen towards the centers of low redshift galaxies (Schmeltz et. al. 1986). Under the assumption that the absorption arises from a rotating galaxy disk, the large OH velocity width implies that the covering factor of the OH absorbing gas is $\sim 1$. In the Nair et. al. model, the distances of the SW and NE images from the lens center are 1.8~kpc and 3.8~kpc respectively (for $H_0 = 75$~km/s/Mpc and $q_0 = 0.5$), and hence OH would have to be widespread in the central 5--6 kpc of the galaxy. The HI spectrum is highly asymmetric, with a peak at $-148$~km/s, the same velocity where weak molecular absorption is seen against the NE core. The HI peak is then presumably gas seen in absorption against the very compact NE component. If one assumes that this component has $\sim 1/3$ of the total flux then for the gas lying in front of the NE component $N_H/T_{sys} \sim 10^{19}$, compatible with galactic numbers of $N_H \sim 10^{21}$~cm$^{-2}$, and $T_{sys} \sim 100$~K. The red wing of the HI absorption profile shows a weak but resolved feature at zero velocity, corresponding to the deep molecular absorption. The contrast in the ratio of HI optical depth at the two velocities, compared to that of the OH molecules, is striking. One possibility is that the gas in front of the SW component is primarily molecular, (i.e. similar to what is seen in many early type spirals). Because the size of the radio source at 753 MHz is estimated to be at least 200 milliarcseconds (cf Patnaik and Porcas, 1995) corresponding about 1.5 kpc, several orders of magnitude larger than at mm wavelengths, this lack of HI must indicate a genuine lack of HI in a substantial part of the inner galaxy. This, in conjunction with the OH spectrum, then suggests that the $z=0.89$ system is an early type spiral with a large central molecular disk, at least 5--6~kpc in size. The broad component in the HI spectrum is presumably the result of HI seen in absorption primarily against the steep spectrum ring. Since at low frequencies the ring has no gaps and the center of the lensing galaxy must lie inside the ring, then without recourse to any specific lensing model it follows that the ring must cut across both the receding and approaching sides of the major axis (Figure~3). For reasonable rotation curves the ring will cut across the major axis well beyond the rising part of the rotation curve. In principle then, the velocity width of the HI spectrum ($\sim 260$~ km/s) corresponds to twice the rotation velocity of the galaxy (apart from an inclination correction). In practice however since the emission from the ring is weakest at the points where it cuts across the major axis, the rotation velocity could be somewhat underestimated. From the model in Nair et. al. (1993) it is straighforward to compute the velocity that one should see in absorption against the SW and the NE cores. The observed velocities are indeed obtained provided one changes the position angle slightly. The inclination angle in the Nair et. al. model is $\sim 40\degr$, however this gives the mass inside the central 4~kpc as $\sim 4\times 10^{10}~M\sun$, which is somewhat low for a source redshift $\sim 1.5-2$. As suggested by Wiklind \& Combes (1998), the inclination angle may be closer to $\sim 20\degr$, and the true rotation velocity more like $\sim 300$~km/s, more typical of early type spirals. In summary then, the OH and HI spectra are consistent with the lens being an early type spiral at a redshift of $z\sim 0.89$.
\section{Introduction} QED is the most successful quantum field theory in the phenomenological point of view. In fact, QED perfectly describes the electromagnetic interaction at low energies. However, there is a question whether QED is a fully consistent theory beyond the perturbation theory. QED has the Landau ghost problem\cite{landau} and the renormalized coupling constant vanishes. This means QED may be trivial as a quantum field theory, and it can only be regarded as a low energy effective theory. On the other hand, Miransky suggested that QED is non-trivial\cite{Miransky}. He investigated a truncated Schwinger-Dyson equation for the fermion propagator and found a continuous chiral phase transition. He claimed that the chiral symmetry is spontaneously broken in the strong coupling phase. After his work, non-perturbative studies of QED have been done extensively\cite{leung,Kondo,yamawaki}. Some of numerical simulations were carried out to understand whether QED is trivial or not. Kogut et al. claimed that the existence of a chiral phase transition was confirmed by numerical studies\cite{kogut}. On the other hand, DESY-J{\"u}lich group claimed that QED is a trivial theory which is described by a Gaussian fixed point, and the critical behavior around it is similar to the one of the $\lambda\phi^4$ model\cite{DESY}. This controversy is not resolved yet. Recently, there has been much progress in the understanding of the non-perturbative dynamics of {\it N}=1 and {\it N}=2 supersymmetric four-dimension field theories. The exact superpotential can be derived in {\it N}=1 supersymmetric QCD (SQCD: supersymmetric SU$(N_c)$ gauge theory with vector-like matters) \cite{Seiberg}, and the models with various gauge symmetries and matter contents have been investigated. Seiberg and Witten derived the exact low energy effective action for {\it N}=2 supersymmetric SU(2) Yang-Mills theory in Coulomb phase up to two derivatives\cite{s-w1}, and generalized it to the case of {\it N}=2 SQCD \cite{s-w2}. Their method was applied to the different gauge groups and the solution was obtained. Since we can derive the exact low energy effective action (LEEA) of {\it N}=2 supersymmetric gauge theories, we can expect to extract the exact information of non-supersymmetric gauge theories, QED and QCD, for example. A simple way to break supersymmetry is to add soft supersymmetry breaking terms. In Refs. \cite{masiero,peskin,sakai,hsu,martin,marino} soft breaking terms are used to explore {\it N}=1 supersymmetric QCD and the phase structure of these theories in the absence of supersymmetry. We will focus on {\it N}=2 supersymmetric QED (SQED) with a massless matter to explore {\it N}=0 QED. It is well known that Fayet-Iliopoulos (FI) term spontaneously breaks supersymmetry in {\it N}=2 SQED \cite{fayet,tree,Ivanov}. We construct the exact LEEA of SQED with FI term not introducing the soft breaking terms by hand. In Refs. \cite{s-w1,s-w2}\cite{masiero,peskin,sakai,hsu,martin,marino} {\it N}=1 superfields were used to describe {\it N}=2 supersymmetric theories. Therefore, {\it N}=2 supersymmetry was not manifest in those works. We can use constrained superfields on the standard {\it N}=2 superspace \cite{grimm}, but these are not appropriate to the construction and the analysis of the LEEA, because the description becomes extremely complicated when the interaction is included on. An elegant off-shell formulation of {\it N}=2 supersymmetry is the harmonic superspace formalism, developed by Galperin et al. \cite{Galperin}. In this formalism superfileds are unconstrained and we do not need to solve complicated constraints. {\it N}=2 supersymmetry is manifest at each step of the calculation. We will see that this formalism is very powerful for constructing the LEEA in this paper. The plan of this paper is as follows. In Sec.II we briefly review the harmonic superspace formalism stressing some important points for our main. In Sec.III we construct the LEEA of SQED without the FI term as the first step. In Sec.IV we extend the discussion in the previous section to the case with FI term. In Sec.V we analyze the effective potential of the LEEA which is obtained in the previous section, and discuss the vacuum structure of N=0 QED. Sec.VI is devoted to the conclusion. Our notations and conventions are summarized in Appendix. \newcommand{\theta^+}{\theta^+} \newcommand{{\bar{\theta}}^+}{{\bar{\theta}}^+} \section{Harmonic Superspace Formalism} We briefly review some of the basics of the harmonic superspace formalism (HSS). HSS is the formalism for {\it N}=2 extended supersymmetry developed by Galperin et al. \cite{Galperin}. The standard {\it N}=2 superspace is parameterized by the coordinates \begin{eqnarray} \{x^\mu, \theta_{\alpha i},{\bar{\theta}}_{\dot{\alpha}}^i \}, \end{eqnarray} where $\alpha$ is the spinor index and $i$ is SU(2$)_{{\rm R}}$ index. The key ingredient in HSS is the harmonic variables $u_i^\pm$ which parameterize the coset space SU(2$)_{{\rm R}}$/U(1). The variables satisfy the relation \begin{equation} u^{+i}u_i^-=1, \end{equation} where $\pm$ denote U(1) charge $\pm 1$. The variables of the harmonic superspace in the central basis (CB) are \begin{equation} {\bf CB}:\{{x^\mu,\theta_{\alpha i},{\bar{\theta}}_{\dot{\alpha}}^i,u_i^\pm}\}. \end{equation} Harmonic superfields are the functions of these variables. In CB the differentiation by the harmonic variables are defined as \begin{eqnarray} &D^{++}=u^{+i}\displaystyle\frac{\partial}{\partial u^{-i}},\hspace{2mm} D^{--}=u^{-i}\displaystyle\frac{\partial}{\partial u^{+i}},& \end{eqnarray} and the integration over $u$ is defined by the following rules: \begin{eqnarray} &\displaystyle\int du\hspace{1mm}1 = 1,&\\ &\displaystyle\int du\hspace{1mm}u_{(i_1}^+\cdots u_{i_n}^+u_{j_1}^-\cdots u_{j_m)}^-=0,\hspace{3mm}n+m>0, \end{eqnarray} where the parenthesis mean symmetrization of SU(2$)_{{\rm R}}$ indices. Namely, the $u$ integration is defined to pick up the SU(2$)_{{\rm R}}$ singlet part. The Lagrangian which is described by the harmonic superfields is not manifestly real under the usual complex conjugation. However, it is real under the conjugation which is the combination of the usual complex conjugation and the star conjugation. The star conjugation for the harmonic variables are defined by \begin{eqnarray} (u_i^+)^*=u_i^-,\hspace{2mm}(u_i^-)^*=-u_i^+, \end{eqnarray} and other quantities are singlet under the conjugation. The harmonic variables are transformed under the combined conjugation as \begin{eqnarray} \displaystyle {\overline{u^{\pm i}}}^*=-u_i^\pm,\hspace{2mm}{\overline{u_i^\pm}}^*=u^{\pm i}. \end{eqnarray} There is another important basis called the analytic basis (AB): \begin{eqnarray} {\bf AB}&:&\{x^\mu_A,\theta_\alpha^\pm,{\bar{\theta}}_{\dot{\alpha}}^\pm,u_i^\pm\},\\ &&x^\mu_A=x^\mu-2i\theta^{(i}\sigma^\mu{\bar{\theta}^{j)}}u_i^+u_j^-, \nonumber\\ &&\theta^\pm=\theta^i u_i^\pm,\hspace{3mm}{\bar{\theta}}^\pm={\bar{\theta}}^i u_i^\pm. \nonumber \end{eqnarray} Irreducible harmonic superfields are not the function of the entire variables of AB or CB but the function on their subspaces, the analytic subspace (ASS) or the chiral subspace (CSS). ASS is defined by \begin{eqnarray} {\bf ASS}:\zeta_A&=&\{x^\mu_A,\theta^+_\alpha,{\bar{\theta}}^+_{\dot{\alpha}},u^\pm_i\}, \end{eqnarray} and it is an invariant subspace under {\it N}=2 supersymmetry transformation. This fact allows one to define the analytic superfields which satisfy the analyticity conditions \begin{eqnarray} D^+\phi^{(q)}(\zeta_A)={\bar{D}}^+\phi^{(q)}(\zeta_A)=0, \end{eqnarray} where \begin{eqnarray} D^+=D^i u_i^+=\displaystyle\frac{\partial}{\partial \theta^-},\hspace{2mm} {\bar{D}}^+={\bar{D}}^i u_i^+=\frac{\partial}{\partial {\bar{\theta}}^-}, \end{eqnarray} and $q$ denotes U(1) charge of the field. There are two basic supermultiplets in the {\it N}=2 supersymmetry: the hypermultiplet and the vectormultiplet. Fayet-Sohnius(FS) superfield \cite{fs} describes the complex hypermultiplet whose on-shell physical components are $(f^i,\psi,\varphi)$, where $f^i$ is a complex scalar in SU(2$)_{{\rm R}}$ doublet, and $(\psi$, $\varphi)$ is a Dirac spinor\footnote{There is another harmonic superfield, Howe-Stelle-Townsend superfield, which describes real hypermultiplet\cite{hst}.}. The superfield with U(1) charge $+1$ is written down as \begin{eqnarray} \phi^{+}(\zeta_A)&=&F^{+}(x_A,u^\pm)+\sqrt{2}\theta^+\psi(x_A,u^\pm) +\sqrt{2}{{\bar{\theta}}^+}{\bar{\varphi}}(x_A,u^\pm)\nonumber\\ &+&\theta^+\ptheta M^{-}(x_A,u^\pm)+{\bar{\theta}}^+\bptheta N^{-}(x_A,u^\pm)\nonumber \\ &+&\theta^+\sigma^\mu{\bar{\theta}}^+ V_\mu(x_A,u^\pm) +\sqrt{2}{\bar{\theta}}^+\bptheta\theta^+\xi_\alpha^{--}(x_A,u^\pm)\nonumber \\ &+&\sqrt{2}{\bar{\theta}}^+{\bar{\chi}}^{--}(x_A,u^\pm) +\theta^+\ptheta{\bar{\theta}}^+\bptheta D^{---}(x_A,u^\pm), \end{eqnarray} where $F^+,M^-,N^-$ and $D^{---}$ are complex scalar fields, $\psi,{\bar{\varphi}},\xi^{--}$ and ${\bar{\chi}}^{--}$ are Weyl fermion fields and $V_\mu$ is a complex vector field. Each component field can be expanded in $u^\pm_i$. For example, \begin{eqnarray} F^{+}(x_A,u^\pm)=\displaystyle\sum_{n=0}^\infty f^{(i_1\cdots i_{n+1}j_1\cdots j_n)}(x_A)u^+_{(i_1}\cdots u^+_{i_{n+1}}u^-_{j_1}\cdots u^-_{j_n)}. \end{eqnarray} Therefore, FS superfield includes infinite number of auxiliary fields. The action for a free complex FS hypermultiplet is given by \begin{eqnarray} S_{{\rm FS}}=\displaystyle\int d\zeta_A^{(-4)}du \hspace{1mm}{\overline{\phi^+}}^*D^{++}\phi^+,\label{eq:hm} \end{eqnarray} where $D^{++}$ is a covariant derivative in AB given by Eq. (\ref{eq:cov}) and $d\zeta_A^{(-4)}$ is the analytic measure defined by \begin{eqnarray} d\zeta_A^{(-4)}=d^4 x_A d^2\theta^+ d^2{\bar{\theta}}^+ . \end{eqnarray} Solving the equation of motion $D^{++}\phi^+=0$, we can easily check that only the physical components $(f^i,\psi,\varphi)$ remain and follow free equation of motions. The on-shell physical components of a vectormultiplet are $(A,A_\mu,\lambda^i)$, where $A$ is a complex scalar, $A_\mu$ is a vector field and $\lambda^i$ is a Majorana spinor in SU(2$)_{{\rm R}}$ doublet. A vectormultiplet is described by the dimensionless analytic superfield $V^{++}$ of U(1) charge +2. It transforms under gauge transformation as \begin{eqnarray} \delta V^{++}=-D^{++}\lambda(x_A,u^\pm)\label{eq:gtrans} \end{eqnarray} in abelian case, where $\lambda$ is an analytic superfield with U(1) charge $0$. Here $V^{++}$ is chosen to be real, namely, \begin{eqnarray} V^{++}={\overline{V^{++}}}^*. \end{eqnarray} If we take the Wess-Zumino-like gauge, \begin{eqnarray} V^{++}(\zeta_A)&=&\theta^+\ptheta\frac{1}{\sqrt{2}}A(x_A,u^\pm)+{\bar{\theta}}^+\bptheta\frac{1}{\sqrt{2}}A^*(x_A,u^\pm)+i\theta^+\sigma^\mu{\bar{\theta}}^+ A_\mu(x_A,u^\pm) \nonumber\\ &&+{\bar{\theta}}^+\bptheta\theta^+ 2\lambda^-(x_A,u^\pm) +\theta^+\ptheta{\bar{\theta}}^+ 2{\bar{\lambda}}^-(x_A,u^\pm) +\theta^+\ptheta{\bar{\theta}}^+\bptheta D^{--}(x_A,u^\pm), \end{eqnarray} where $D^{--}=D^{(ij)}u_i^- u_j^-$ is a real auxiliary field in SU(2$)_{{\rm R}}$ triplet. CSS is defined by \begin{eqnarray} {\bf CSS}:\zeta_R&=&\{x^\mu_R,\theta_{\alpha i},u_i^\pm\},\\ &&x^\mu_R=x^\mu-i\theta^i\sigma^\mu{\bar{\theta}}_i, \nonumber \end{eqnarray} and the gauge field strength superfield $W$ is described as a function on it. \begin{eqnarray} W(x_{R},\theta)&=&-\displaystyle\frac{1}{4}\int du ({\bar{D}}^+)^2 V^{++}\\ &=&\displaystyle\frac{1}{\sqrt{2}}A^*(x_R) -\frac{1}{3\sqrt{2}}\epsilon_{ik}\epsilon_{jl}(\theta^i\theta^j)(\theta^k\theta^l)\Box A(x_R) -\frac{1}{4}\theta^i\sigma^\mu{\bar{\sigma}}^\nu\theta_i F_{\mu\nu}(x_R) \nonumber \\ &&+\theta^i\lambda_i(x_R)+\frac{2}{3}i(\theta^i\theta^j)(\theta^k\sigma^\mu\partial_\mu{\bar{\lambda}}^l(x_R))\epsilon_{ik}\epsilon_{jk}+\frac{1}{3}\epsilon_{ik}\epsilon_{jk}(\theta^i\theta^j)D^{(kl)}(x_R).\label{eq:fieldst} \end{eqnarray} The superfield which is a function on CSS is called the chiral superfield. Note that the chiral superfield $W$ does not explicitly depend on ${\bar{\theta}}_i$ and $u_i^\pm$ and can not the function on ASS. Also note that the analytic superfield can not be described as a function on CSS. The action of a vectormultiplet is given by \begin{eqnarray} S_{{\rm gauge}}=\frac{1}{4\pi}{\rm Im}\tau\int d\zeta_R W^2, \end{eqnarray} where $\tau=i\frac{4\pi}{e^2}+\frac{\Theta}{2\pi}$ with that $e$ is the gauge coupling and $\Theta$ is the vacuum angle. $d\zeta_R=d^4x_R d^4\theta du$ is the chiral subspace measure. We write down the tree-level action of SQED with single matter as \begin{eqnarray} S=\displaystyle\int d\zeta_A^{(-4)}du\hspace{1mm}{{\overline{\phi^+}}}^*(D^{++}+2iV^{++}) \phi^+ + \frac{1}{4\pi}{\rm Im}\tau\int d\zeta_R W^2. \end{eqnarray} The integrand of the first term (the analytic part) must have U(1) charge $+4$ and does not explicitly depend on $\theta^-$ and ${\bar{\theta}}^-$, i.e., it must be analytic. The chiral superfield does not appear in the analytic part, because the chiral superfield does not satisfy the analyticity. Similarly, we find that the analytic superfield does not appear in the integrand of the second term (the chiral part). The chiral part does not explicitly depend on ${\bar{\theta}_i}$ and $u_i^\pm$. These facts are important for constructing the LEEA. \section{Construction of the LEEA without Fayet-Iliopoulos term}\label{leea1} In this section we construct the LEEA of SQED with single massless matter using the harmonic superspace formalism. In the next section we apply the method used in this section to the case of including FI term. The tree-level action leads the scalar potential \begin{eqnarray} V=|\sqrt{2}A|^2{\bar{f}}^i f_i+\displaystyle\frac{e_0^2}{2}({\bar{f}}^i f_i)^2, \end{eqnarray} where $e_0$ is the bare coupling constant, $A$ and $f^i$ are the complex scalar fields in the vectormultiplet and FS hypermultiplet, respectively. The classical moduli space is parameterized by the vacuum expectation value of the complex scalar field $A$. In case of single matter, $f^i$ has no vacuum expectation value and the gauge symmetry is not broken. Namely, the theory is always in Coulomb phase. If we consider multiple matter, the moduli space has Higgs branch in which the gauge symmetry is broken. Our strategy of getting the LEEA is the same which was developed in Ref. \cite{Seiberg}. The LEEA must be invariant under the enlarged symmetry transformation in which the parameters of the theory transform. These parameters can be considered as the vacuum expectation values of some external superfields. The holomorphy (or analyticity) also constrains the LEEA. By using the information obtained in the weak coupling limit, we can determine the LEEA. The transformation laws of the fields and parameters in the fundamental theory is summarized in table \ref{sym1}. We assume that there is no confiment at low energies. If the resultant LEEA has no inconsistency, we can conclude that this assumption is justified. The general form of the LEEA of the chiral part (lowest order in the derivative expansion) is given by \begin{eqnarray} {\cal L}_{{\rm C}}=\displaystyle\frac{1}{4\pi}{\rm Im}\int d^4\theta g(W,\Lambda),\label{eq:gauge1} \end{eqnarray} where $g(W,\Lambda)$ is a holomorphic function which satisfies the following conditions. \begin{enumerate} \item U(1) charge $0$. \item mass dimension $2$. \item U(1$)_{{\rm R}}$ charge 4. \item gauge singlet. \end{enumerate} We stress again that the FS superfield can not appear in the chiral part. The parameter $\Lambda$ can be understood as the vacuum expectation value of the lowest component of a chiral superfield. The above conditions restrict Eq. (\ref{eq:gauge1}) to be the form \begin{eqnarray} {\cal L}_{{\rm C}}=\displaystyle\frac{1}{4\pi}{\rm Im}\int d^4\theta G\left(\frac{\Lambda}{W}\right)W^2. \end{eqnarray} We can estimate $G$ at one-loop level in the weak coupling limit $\Lambda\rightarrow \infty$. Namely, we can get \begin{eqnarray} \displaystyle\lim_{\Lambda\rightarrow \infty}G\left(\frac{\Lambda}{W}\right)=\frac{i}{\pi}\ln\frac{\Lambda}{W}. \end{eqnarray} Thus we obtain \begin{eqnarray} G\left(\frac{\Lambda}{W}\right)=\frac{i}{\pi}\ln\frac{\Lambda}{W}+\tilde{G}\left(\frac{\Lambda}{W}\right), \end{eqnarray} where $\tilde{G}$ includes the non-perturbative effect. We assume that $\tilde{G}$ does not have singularities, namely, all massless particles have been already included. Then, the Liouville theorem leads \begin{eqnarray} \tilde{G}\left(\frac{\Lambda}{W}\right)={\rm constant}. \end{eqnarray} Therefore, the chiral part is determined as \begin{eqnarray} {\cal L}_{{\rm C}}=\displaystyle\frac{1}{4\pi}{\rm Im}\int d^4\theta \frac{i}{\pi}W^2\ln\frac{\Lambda}{W}.\label{eq:gauge} \end{eqnarray} This is exactly the same result given by Seiberg and Witten \cite{s-w2}. Note that the singularity at $\langle W \rangle=0$ is not removed in spite of considering the elementary matter field. The theory is not defined at $\langle W \rangle=\langle A \rangle=0$ within our assumptions. Next, we determine the LEEA of the analytic part. The general form is given by \begin{eqnarray} {\cal L}_{{\rm A}}=\displaystyle\int d^2\theta^+ d^2{\bar{\theta}}^+ du \hspace{1mm}f^{(+4)}(\phi^+,{\overline{\phi^+}}^*,V^{++},{\cal D^{++}}), \end{eqnarray} where ${\cal D^{++}}$ represents the covariant derivative ${\cal D^{++}}=D^{++}+2iV^{++}$. Analytic function $f^{(+4)}$ must satisfy the following conditions. \begin{enumerate} \item U(1) charge 4. \item mass dimension 2. \item U(1$)_{\rm R}$ charge 0. \item gauge singlet. \end{enumerate} We stress again that the chiral superfield $W$ can not appear in the analytic part. Considering the above conditions, we obtain \begin{eqnarray} {\cal L}_{{\rm FS}}=\displaystyle\int d^2\theta^+ d^2{\bar{\theta}}^+ du \hspace{1mm}{\overline{\phi^+}}^*{\cal D^{++}}\phi^+. \end{eqnarray} Surprisingly, this is the same form of the tree-level one. The first derivation of the LEEA of the hypermultiplet of SQED and SQCD was done in Ref. \cite{ketov} using the harmonic superspace formalism. In Ref. \cite{ketov} the self-interaction of the massive FS hypermultiplet is derived by the perturbative calculation: \begin{eqnarray} \Delta{\cal L}=\lambda\displaystyle\int d^2\theta^+ d^2{\bar{\theta}}^+ du \hspace{1mm} ({\overline{\phi^+}}^*\phi^+)^2,\end{eqnarray} where $\lambda$ includes an infrared cutoff. The self-interaction term does not appear in our method based on the symmetry and holomorphy even in the massive case. It is expected that the infrared divergence disappears by summing up all the one-loop diagrams with external FS superfields, and only the higher order terms in the derivative expansion are obtained. The total LEEA of SQED is \begin{eqnarray} S_{{\rm eff}}=\displaystyle\int d\zeta_{A}^{(-4)} du \hspace{1mm}{\overline{\phi^+}}^*{\cal D^{++}}\phi^++\displaystyle\frac{1}{4\pi}{\rm Im}\int d\zeta_R \frac{i}{\pi}W^2\ln\frac{\Lambda}{W}. \end{eqnarray} We remark the modification of the moduli space by the quantum effect. The quantum effect forbids a part of the moduli space $\langle W \rangle=\langle A \rangle >\Lambda$ where the effective coupling $\alpha_{{\rm eff}}$ is negative. \section{Construction of the LEEA with Fayet-Iliopoulos term} We construct the LEEA of SQED with spontaneous supersymmetry breaking to get some exact results of N=0 QED. In case of SQED, we can introduce FI term \begin{eqnarray} {\cal L_{{\rm FI}}}=\displaystyle\int d^2\theta^+ d^2{\bar{\theta}}^+ du \hspace{1mm}\xi^{++}V^{++}=\frac{1}{3}\xi^{ij}D_{(ij)}, \hspace{2mm}\xi^{++}\equiv\xi^{ij}u_i^+u_j^+ \label{eq:F.I} \end{eqnarray} to break supersymmetry spontaneously, where $\xi^{ij}$ includes three real parameters $\xi^{(a)}$ of mass dimension 2: \begin{eqnarray} \xi&=&i\xi^{(a)}(\sigma^{a}\epsilon)= \left( \begin{array}{cc} i\xi^{(1)}+\xi^{(2)} & -i\xi^{(3)} \\ -i\xi^{(3)} & -i\xi^{(1)}+\xi^{(2)} \end{array} \right). \label{eq:xi} \end{eqnarray} The procedure of constructing the LEEA is the same as that in the previous section. The transformation laws for the fields and parameters are summarized in table \ref{sym2}. The parameters $\xi^{ij}$ can be understood as the vacuum expectation value of the analytic superfield $\xi^{++}$. First we consider the LEEA of the chiral part. Repeating the same arguments in the previous section, we obtain the general form \begin{eqnarray} {\cal L}_{{\rm C}}=\displaystyle\frac{1}{4\pi}{\rm Im}\int d^4\theta g(W,\Lambda)=\frac{1}{4\pi}{\rm Im}\int d^4\theta G\left(\frac{\Lambda}{W}\right)W^2. \end{eqnarray} This is exactly the same form that is obtained in the case without FI term. The coefficient $\xi^{++}$ can not be included in $G$. After all, using the one-loop result for $G$, the LEEA of the chiral part is given by Eq. (\ref{eq:gauge}). Next we consider the LEEA of the analytic part. The general form is \begin{eqnarray} {\cal L}_{{\rm A}}=\displaystyle\int d^2\theta^+ d^2{\bar{\theta}}^+ du \hspace{1mm}f\left(\frac{\xi^{++}}{\phi^+{\overline{\phi^+}}^*}\right){\overline{\phi^+}}^*{\cal D^{++}}\phi^+. \end{eqnarray} We can estimate the function $f\left(\frac{\xi^{++}}{\phi^+{\overline{\phi^+}}^*}\right)$ in the weak coupling limit $\xi^{++}\rightarrow 0$ using the perturbation theory. We find that there is no one particle irreducible diagram which includes $\xi^{++}$ and conclude \begin{eqnarray} \displaystyle\lim_{\xi\rightarrow 0} f\left(\frac{\xi^{++}}{\phi^+{\overline{\phi^+}}^*}\right)={\rm constant}. \end{eqnarray} We can make the constant unity by rescaling the field $\phi^+$. Including the non-perturbative effect, $f$ is given by \begin{eqnarray} f\left(\frac{\xi^{++}}{\phi^+{\overline{\phi^+}}^*}\right)=1+\tilde{f}\left(\frac{\xi^{++}}{\phi^+{\overline{\phi^+}}^*}\right), \end{eqnarray} where $\tilde{f}$ describes non-perturbative effect. Here, we assume again that all massless fields have been already included and the analytic function has no singularity. The Liouville theorem leads \begin{eqnarray} f\left(\frac{\xi^{++}}{\phi^+{\overline{\phi^+}}^*}\right)=1. \end{eqnarray} Therefore, after all, the LEEA of the analytic part is exactly the same with that is obtained in the case without FI term. The FI term in Eq. (\ref{eq:F.I}) is the exact form. An analytic function $h\left(\frac{\xi^{++}}{\phi^+{\overline{\phi^+}}^*}\right)$ seems to be allowed as the coefficient function of FI term. However the function must be a constant due to the gauge invariance. Note that $V^{++}$ is gauge invariant up to the total derivative. We conclude that the LEEA of SQED with FI term is given by \begin{eqnarray} S_{{\rm eff}}=\displaystyle\int d \zeta_A^{(-4)} du \hspace{1mm}{\overline{\phi^+}}^*{\cal D^{++}}\phi^++\displaystyle\frac{1}{4\pi}{\rm Im}\int d\zeta_R \frac{i}{\pi}W^2\ln\frac{\Lambda}{W}+\displaystyle\int d\zeta_A^{(-4)} du \hspace{1mm}\xi^{++}V^{++}. \end{eqnarray} \section{Potential analysis of N=0 QED }\label{potesec} In this section, we write down and analyze the effective potential of the LEEA which is obtained in the previous section. We take the polar decomposition \begin{eqnarray} A=a e^{i\sigma}, \end{eqnarray} where $a$ and $\sigma$ are real scalar fields. The contribution to the potential from the analytic part including FI term is \begin{eqnarray} V_{{\rm A}}=(\sqrt{2}a)^2{\bar{f}}^i f_i+\displaystyle\frac{2}{3}i{\bar{f}}^i f^j D_{(ij)}-\frac{1}{3}\xi^{ij}D_{(ij)}. \end{eqnarray} The contribution from the chiral part is \begin{eqnarray} V_{{\rm C}}=\displaystyle\frac{1}{8\pi^2}\left(\frac{1}{9}D^{(ij)}D_{(ij)}\ln\frac{\Lambda}{a}-\frac{1}{6}D^{(ij)}D_{(ij)}+{\rm h.c}\right). \end{eqnarray} Using the equation of motion of the auxiliary field $D^{(ij)}$, we obtain the total scalar potential as \begin{eqnarray} V_{{\rm eff}}&=&V_{{\rm A}}+V_{{\rm C}}\nonumber\\ &=&(\sqrt{2}a)^2{\bar{f}}^i f_i-\displaystyle\frac{4\pi^2}{\ln\frac{\Lambda^\prime}{a}}\left({\bar{f}}^if^j+\frac{i}{2}\xi^{ij}\right)\left({\bar{f}}_if_j+\frac{i}{2}\xi_{ij}\right), \end{eqnarray} where $\Lambda^\prime=\Lambda e^{-3/2}$. Note that the potential is independent of the scalar field $\sigma$. The vacuum expectation value of $\sigma$ is unphysical, since $\Theta$ term in U(1) gauge theory has no meaning. The extremal conditions for $a$ and $f$ are \begin{eqnarray} \displaystyle\frac{\partial V_{{\rm eff}}}{\partial {\bar{f}}^i}&=&\left\{(\sqrt{2}a)^2\epsilon_{ij}-\displaystyle\frac{8\pi^2}{\ln\frac{\Lambda^\prime}{a}}\left({\bar{f}}_if_j+\frac{i}{2}\xi_{ij}\right)\right\}f^j=0,\\ \displaystyle\frac{\partial V_{{\rm eff}}}{\partial a}&=&2a{\bar{f}}^if_i+\frac{1}{a}\displaystyle\frac{4\pi^2}{(\ln\frac{\Lambda^\prime}{a})^2}\left({\bar{f}}^if^j+\frac{i}{2}\xi^{ij}\right)\left({\bar{f}}_if_j+\frac{i}{2}\xi_{ij}\right)=0. \end{eqnarray} The solution is \begin{eqnarray} f^1&=&f^2=0,\nonumber\\ a&\rightarrow&\infty. \end{eqnarray} This solution gives $V_{{\rm eff}}=0$ and {\it N}=2 supersymmetry seems to be unbroken. However, such a solution is ruled out, since $a>\Lambda$ is not allowed by the quantum deformation of the moduli space. Therefore we conclude that there is no stable vacuum in the LEEA of SQED with FI term under the assumption of no confinement\footnote{Without FI term there is a stable vacuum, of course. We can define a theory on a point in the moduli space except for a=0 and $a>\Lambda$. For small $a$, the LEEA reduces to the one obtained in the perturbation theory.} . Here, we summarize how the moduli space has been deformed. In the classical theory the moduli space is parameterized by $a$, and any value of $a$ is possible (fig.\ref{clas}(a)). By the quantum effect the region $a>\Lambda$ and a point $a=0$ are forbidden (fig.\ref{clas}(b)). By including FI term remaining moduli space is lifted up and slopes down to $a=0$ axis, and no stable vacuum exists (fig.\ref{clas}(c)). We interpret this results as follows. Recall that we assume that the confinement does not occur at low energies. Thus, the result no stable vacuum in the LEEA of SQED with FI term, suggests that the confinement may occur at low energies. If we assume confinement at low energies, we may be able to remove the singularity at $a=0$ and may obtain a stable vacuum. The shape of the scalar potential is given in fig.\ref{rittai}. The vacuum energy incorrectly takes negative value for $a>\Lambda^\prime$. For $a<\Lambda^\prime$, the potential slopes down to $a=0$ axis where the theory can not be defined. The slice of the potential along $f=0$ axis is shown in fig.\ref{danmen}. We have almost the same form as in fig.\ref{danmen} for any slice along $f\neq 0$. The structure along the axis of the constant $a$ is a little complicated. We can understand it by referring the masses of two fields $f^i$. They are obtained as \begin{eqnarray} m^2_{f_1,f_2}=(\sqrt{2}a)^2\pm\displaystyle\frac{2\pi^2}{\ln\frac{\Lambda}{a}}\left\{\sum_{a=1}^3(\xi^{(a)})^2\right\}^{\frac{1}{2}}. \end{eqnarray} Note that one of the squared masses can become negative for small value of $a$ satisfying condition \begin{eqnarray} \displaystyle 2\pi^2\left\{\sum_{a=1}^3(\xi^{(a)})^2\right\}^{\frac{1}{2}}>(\sqrt{2}a)^2 \ln\frac{\Lambda^\prime}{a}.\label{eq:cond} \end{eqnarray} Fig.\ref{danmen2} shows the typical shape of the slice along $a\neq 0$ for small $a$. \section{Conclusion} To obtain some exact results of QED, we constructed the LEEA of SQED with single massless matter including FI term. We assumed that the confinement does not occur at low energies and the LEEA is described by elementary fields. We found that the harmonic superspace formalism is very useful for applying symmetry and holomorphy in the construction. We reproduced the LEEA of the chiral part which is coincide with the result given by Seiberg and Witten. We constructed the LEEA of the analytic part including FI term. This part was the tree-level exact. We wrote down the scalar potential of the LEEA and analyzed it. We found that there is no stable vacuum, and could not define the theory. We interpret this result as an evidence of the confinement at low energies in non-supersymmetric QED. If we assume there is confinement at low energies, we may get rid of the singularity at $a=0$ and obtain a stable vacuum.
\section{INTRODUCTION} The question of what happens to a compact object that is fed mass at rates far higher than its Eddington limit has a long history (Shakura \& Sunyaev, 1973; Kafka \& M\'esz\'aros, 1976; Begelman, 1979). In the context of accreting binary systems, this problem is particularly acute because of the possibility of common--envelope (CE) evolution at such rates. That is, the accreting component may be unable either to accept or to expel the mass at a sufficiently high rate to avoid the formation of an envelope engulfing the entire binary system. The frictional drag of this envelope can shrink the binary orbit drastically. If the resulting release of orbital energy is enough to unbind the envelope, the binary will emerge from the common envelope with a smaller separation; if not, the binary components may coalesce. CE evolution is probably required for the formation of binaries such as cataclysmic variables, in which the binary separation is far smaller than the radius of the accreting white dwarf's red giant progenitor. However, it is in general an open question whether CE evolution occurs in any given binary. This question is thrown into sharp relief by recent work on the evolution of the low--mass X--ray binary Cygnus X--2 (King \& Ritter, 1999), which has a period of 9.84~d. The rather precise spectroscopic information found by Casares, Charles, \& Kuulkers (1998), together with the observed effective temperature of the secondary, shows that this star has a mass definitely below $0.7{\rm M_{\odot}}$ and yet a luminosity of order $150{\rm L_{\odot}}$. King \& Ritter (1999) consider several possible explanations and show that the only viable one is that Cygnus X--2 is a product of early massive Case B evolution. Here `Case B' means that the mass--losing star has finished core hydrogen--burning, and is expanding across the Hertzsprung gap: `early' means that the stellar envelope is radiative rather than convective, and `massive' that the helium core is non--degenerate; see Kippenhahn \& Weigert, 1967. In Cygnus X--2 an initially more massive ($M_{2i} \simeq 3.5{\rm M_{\odot}}$) secondary transferred mass on a thermal timescale ($\sim 10^6$~yr) to the neutron star. This idea gives a satisfying fit to the present observed properties of Cyg X--2, as well as a natural explanation for the large white dwarf companion masses found in several millisecond pulsar binaries with short orbital periods. CE evolution cannot have occurred, as Cyg X--2's long orbital period means that there was far too little orbital energy available for the CE mechanism to have ejected so much mass. Thus an inescapable feature of this picture is that the neutron star is evidently able to eject essentially all of the matter ($\ga 2 - 3{\rm M_{\odot}}$) transferred to it at highly super--Eddington rates $\ga 10^{-6}{\rm M_{\odot}}$~yr$^{-1}$. Indeed, the neutron star mass in Cyg X--2 is rather close to the canonical value of $1.4{\rm M_{\odot}}$. The aim of this paper is to determine under what conditions such expulsion can occur without the system going into a common envelope. \section{EXPULSION BY RADIATION PRESSURE} There are essentially two views as to the fate of matter dumped onto a compact object at a highly super--Eddington rate. In spherically--symmetric, dissipative accretion of an electron--scattering medium, the luminosity generated by infall down to radius $R$ will reach the Eddington limit at a radius \begin{equation} R_{\rm ex} \sim \biggl({\dot M_{\rm tr}\over \dot M_{\rm Edd}}\biggr)R_S, \end{equation} where $\dot M_{\rm tr}$ is the mass infall rate at large radius (i.e. the mass transfer rate from the companion star in our case), $\dot M_{\rm Edd} = L_{\rm Edd}/ c^2$ is the Eddington accretion rate, and $R_S$ is the Schwarzschild radius (Begelman, 1979). This is also the ``trapping radius", below which photon diffusion outward cannot overcome the advection of photons inward. If the compact object is a black hole, the radiation generated in excess of the Eddington limit can thus be swept into the black hole, and lost. If the compact object is a neutron star, however, radiation pressure building up near the star's surface must resist inflow in excess of $\dot M_{\rm Edd}$, causing the stalled envelope to grow outward. This situation would lead to the formation of a common envelope. The outcome may be very different if the accretion flow has even a small amount of angular momentum. Shakura \& Sunyaev (1973) suggested that super--Eddington flow in an accretion disk would lead to the formation of a strong wind perpendicular to the disk surface, which could carry away most of the mass. Such a model (an ``Adiabatic Inflow-Outflow Solution,", or ADIOS) was elaborated by Blandford \& Begelman (1999: hereafter BB99), who considered radiatively inefficient accretion flows in general. BB99 recalled that viscous transfer of angular momentum also entails the transfer of energy outward. If the disk were unable to radiate efficiently (as would be the case at $R < R_{\rm tr}$), the energy deposited in the material well away from the inner boundary would unbind it, leading to the creation of powerful wind. BB99 described a family of self-similar models in which the mass inflow rate decreases inward as $\dot M \propto r^n$ with $0<n<1$. The exact value of $n$ depends on the physical processes depositing energy and angular momentum in the wind. If these are very efficient (e.g., mediated by highly organized magnetic torques) $n$ could be close to zero, in which case little mass would be lost. However, if the wind is produced inefficiently, $n$ would have to be close to 1 and the mass flux reaching the central parts of the accretion disk would be much smaller than the mass transferred from the secondary. For example, two-dimensional hydrodynamical simulations of the evolution of a non-radiative viscous torus (Stone, Pringle \& Begelman 1999) show the development of a convectively driven circulation with little mass reaching the central object, and $n\sim 1$. The development of this strong mass loss is generic and is not related to the assumption of self-similarity. In effect, what is happening is that the energy liberated by a small fraction of the mass reaching the deep gravitional potential serves to unbind the majority of the matter which is weakly bound at large distances. While the specific details of mass loss from super--Eddington flow have not been worked out (in particular, radiation-dominated convection is poorly understood), it is reasonable to assume that the wind will be produced inefficiently, with $n\sim 1$ as in the case of hydrodynamic convection. We also assume that most of the matter will be blown away from $R_{\rm ex}$. Applying equation (1), we find \begin{equation} R_{\rm ex} \simeq 1.3\times 10^{14}\dot m_{\rm tr} \ {\rm cm}, \label{rex2} \end{equation} where $\dot m_{\rm tr}$ is the mass transfer rate expressed in ${\rm M_{\odot}}$ yr$^{-1}$. Note that $R_{\rm ex}$ is independent of the mass of the compact accretor. Since we restrict attention to electron scattering opacity only, we require that hydrogen should be strongly ionized at $R_{\rm ex}$. This is ensured by requiring the radiation temperature near $R_{\rm ex}$ to exceed $T_H \sim 10^4$~K. Since the luminosity emerging from $R_{\rm ex}$ is close to the Eddington limit, we require $L_{\rm Edd} \ga 4\pi R_{\rm ex}^2\sigma T_H^4$, which is satisfied if $R_{\rm ex}/R_S \la 10^7 m_1^{-1/2}$, or equivalently (using equation[1]) \begin{equation} \dot M_{\rm tr}\la 10^7 \dot M_{\rm Edd} m_1^{-1/2} \simeq 2\times 10^{-2} m_1^{1/2} ~{\rm M_{\odot}}\ {\rm yr}^{-1} \label{h} \end{equation} where $m_1 = M_1/{\rm M_{\odot}}$ is the mass of the compact accretor (black hole or neutron star). CE evolution will be avoided if $R_{\rm ex}$ is smaller than the accretor's Roche lobe radius $R_1$. If the accretor is the less massive star (as will generally hold in cases of interest) we can use standard formulae to write \begin{equation} r_1 = 1.9m_1^{1/3}P_{\rm d}^{2/3}, \label{roche} \end{equation} where $r_1 = R_1/{\rm R_{\odot}}$ and $P_{\rm d}$ is the orbital period measured in days. Combining with equation (\ref{rex2}) gives \begin{equation} \dot M_{\rm tr} \la 10^{-3}m_1^{1/3}P_{\rm d}^{2/3}~{\rm M_{\odot}}\ \ {\rm yr}^{-1}. \label{lim1} \end{equation} This form of the limit can be compared directly with observation if we have estimates of the transfer rate, orbital period and the accretor mass. For more systematic study it is useful to replace the dependence on the accretor's Roche lobe by that on its companion's. Thus, since the mass transfer rate is specified by properties of the companion star, which is assumed to fill its Roche lobe radius $R_2$, we eliminate $R_1$ from the condition $R_{\rm ex} \la R_1$ by using the relation \begin{equation} {R_1\over R_2} = \biggl({m_1\over m_2}\biggr)^{0.45}, \label{lobe} \end{equation} (cf King et al., 1997) where $M_2 = m_2{\rm M_{\odot}}$ is the companion mass. Writing $R_2 = r_2{\rm R_{\odot}}$ we finally get the limit \begin{equation} \dot M_{\rm tr} \la 5\times 10^{-4}m_1^{0.45}m_2^{-0.45}r_2~{\rm M_{\odot}}\ {\rm yr}^{-1} \label{lim} \end{equation} on the mass transfer rate if CE evolution is not to occur. \section{AVOIDANCE OF COMMON ENVELOPE EVOLUTION} By specifying the nature of the companion star we fix $m_2, r_2$ and $\dot M_{\rm tr}$ in (\ref{lim}), and so can examine whether CE evolution is likely in any given case. Rapid mass transfer occurs if the companion star is rather more massive than the accretor, since then the act of transferring mass shrinks the donor's Roche lobe. The mass transfer proceeds on a dynamical or thermal timescale depending on whether the donor star's envelope is largely convective or radiative (e.g. Savonije, 1983). In the first case, CE evolution is quite likely to ensue, as the mass transfer rate rises to very high values. However, even in this case it is worth checking the inequality (\ref{lim}) in numerical calculations, as the e--folding time for the mass transfer is $t_e\sim (H/R_2)t_M$, where $H$ is the stellar scaleheight and $t_M$ is the mass transfer timescale set by whatever process (e.g., nuclear evolution) brought the donor into contact with its Roche lobe initially. For main--sequence and evolved stars we have $H/R_2 \sim 10^{-4}, 10^{-2}$ respectively. Thus $t_e$ may be long enough that the companion mass is exhausted before (\ref{lim}) is violated. Thermal--timescale mass transfer is rather gentler, and offers the possibility of avoidance of CE evolution. In addition to the case mentioned above, thermal--timescale mass transfer will also occur if the donor star is crossing the Hertzsprung gap and has not yet developed a convective envelope (i.e., is not close to the Hayashi line), even if it is the less massive star. Detailed calculations (Kolb, 1998) show that in both cases the mass transfer rate is given roughly by \begin{equation} \dot M_{\rm tr} \sim {M_2\over t_{\rm KH}}, \label{th} \end{equation} where \begin{equation} t_{\rm KH} = 3\times 10^7{m_2^2\over r_2l_2}~{\rm yr} \label{kh} \end{equation} was the Kelvin--Helmholtz time of the star when it left the main sequence, and $L_2 = l_2{\rm L_{\odot}}$ was its luminosity. (Note that by definition the donor is not in thermal equilibrium, so an originally main--sequence donor will develop a non--equilibrium structure as mass transfer proceeds.) The condition of a radiative envelope requires a main--sequence mass $m_2 \ga 1$, so we may take \begin{equation} r_2 \sim m_2^{0.8},\ \ l_2 \sim m_2^3. \label{ms} \end{equation} Inserting in (\ref{kh}) and (\ref{th}) we find \begin{equation} \dot M_{\rm tr} \sim 3\times 10^{-8}m_2^{2.8}, \label{trmax} \end{equation} so comparing with (\ref{lim}) we require \begin{equation} m_2 \la 53 m_1^{0.18} \label{mmax} \end{equation} and thus (from \ref{trmax}) \begin{equation} \dot M_{\rm tr,\ max} \sim 2\times 10^{-3}m_1^{0.51}\ {\rm M_{\odot}}\ \ {\rm yr}^{-1}. \label{mtrmax} \end{equation} Hence we expect CE evolution to be avoided in thermal--timescale mass transfer from a main--sequence star, or from a Hertzsprung gap star, provided that it has a radiative envelope. This is in agreement with the assumption of no CE evolution in Cyg X--2 made by King \& Ritter (1999), where the initial donor mass was about $3.5{\rm M_{\odot}}$. \section{CONCLUSIONS} We have derived a general criterion for the avoidance of common--envelope evolution in a binary in which the accretor is a neutron star or a black hole. This shows that thermal--timescale mass transfer from a main--sequence star is unlikely to lead to CE evolution, as is mass transfer from a Hertzsprung gap star, provided that the envelope is radiative. The first possibility allows the early massive Case B evolution inferred by King \& Ritter (1999) for the progenitor of Cyg X--2. SS433 may be an example of the second possibility, with a fairly massive donor star. We will discuss this possibility in detail in a future paper. The considerations of this paper suggest that common--envelope evolution with a neutron--star or black--hole accretor generally requires an evolved donor with a deep convective envelope. This represents a slight restriction on some of the routes invoked in the possible formation of Thorne--$\dot {\rm Z}$ytkow objects. \acknowledgements This research was carried out at the Institute for Theoretical Physics and supported in part by the National Science Foundation under Grant No. PHY94--07194. ARK gratefully acknowledges support by the UK Particle Physics and Astronomy Research Council through a Senior Fellowship. MCB acknowledges support from NSF grant AST95--29170 and a Guggenheim Fellowship. \clearpage
\section{Proem} The interpretation of extensive air shower (EAS) measurements in the PeV domain and above relies strongly on the hadronic interaction model applied when simulating the shower development in the Earth's atmosphere. Such models are needed to describe the interaction processes of the primary particles with the air nuclei and the production of secondary particles. In the EAS Monte Carlo codes the electromagnetic and weak interactions can be calculated with good accuracy. Hadronic interactions, on the other hand, are still uncertain to a large extent. A wealth of data exists on particle production from $p\overline{p}$ colliders up to energies which correspond to 2~PeV/c laboratory momentum and from heavy ion experiments up to energies of 200~GeV/nucleon. However, almost all collider experiments do not register particles emitted in the very forward direction where most of the energy flows. These particles carry the preponderant part of the energy and, therefore, are of utmost importance for the shower development of an EAS. Since most of these particles are produced in interactions with small momentum transfer, QCD is at present not capable of calculating their kinematic parameters. Many phenomenological models have been developed to reproduce the experimental results. Extrapolations to higher energies, to small angles, and to nucleus--nucleus collisions have been performed under different theoretical assumptions. The number of participant nucleons in the latter case is another important parameter which influences the longitudinal development of a shower. Many EAS experiments have used specific models to determine the primary energy and to extract information about the primary mass composition. Experience shows that different models can lead to different results when applied to the same data. Therefore, it is of crucial importance to verify the individual models experimentally as thoroughly as possible. When planning the KASCADE experiment, one of the principal motivations to build the hadron calorimeter was the intention to verify available interaction models by studying the hadronic central core. In the Monte Carlo code CORSIKA \cite{corsika} five different interaction codes have been implemented and placed at the users' disposal. By examining the hadron distribution in the very centre these interaction models are tested. The propagation code itself, viz. hadron transport, decay modes, scattering etc., is checked by looking to the hadron lateral distribution further outside up to distances of 100 m from the core. \section{The apparatus} \begin{figure}\centering \epsfig{file=eps/kalor.ps,width=\textwidth, clip=,bbllx=73,bblly=333,bburx=504,bbury=620} \caption{Sketch of the KASCADE central calorimeter. Detailed view (top) and total view (bottom).} \label{kalor} \end{figure} The KASCADE experiment consists of an array of 252 stations for electron and muon detection and a calorimeter in its centre for hadron detection and spectroscopy. It has been described in detail elsewhere \cite{kascade}. The muon detectors in the array are positioned directly below the scintillators of the electron detectors and are shielded by slabs of lead and iron corresponding to 20 radiation lengths in total. The absorber imposes an energy threshold of about 300 MeV for muon detection. The calorimeter is of the sampling type, the energy being absorbed in an iron stack and sampled in eight layers by ionisation chambers. Its performance is described in detail by Engler et al. \cite{kalorimeter}. A sketch of the set--up is shown in Fig.~\ref{kalor}. The iron slabs are 12--36~cm thick, becoming thicker in deeper parts of the calorimeter. Therefore, the energy resolution does not scale as $1/\sqrt{E}$, but is rather constant varying slowly from $\sigma/E = 20\%$ at 100~GeV to 10\% at 10~TeV. The concrete ceiling of the detector building is the last part of the absorber and the ionisation chamber layer below acts as tail catcher. In total, the calorimeter thickness corresponds to 11 interaction lengths $\lambda_{I}$ for vertical hadrons. On top, a 5~cm lead layer filters off the electromagnetic component to a sufficiently low level. The liquid ionisation chambers use the room temperature liquids tetramethylsilane (TMS) and tetramethylpentane (TMP). A detailed description of their performance can be found elsewhere \cite{prototyp}. Liquid ionisation chambers exhibit a linear signal behaviour with a very large dynamic range. The latter is limited only by the electronics to about $5\times10^{4}$ of the amplifier rms--noise, i.e., the signal of one to more than $10^{4}$ passing muons, equivalent to 10 GeV deposited energy, is read out without saturation. This ensures the energy of individual hadrons to be measured linearly up to 20\,TeV. At this energy, containment losses are at a level of two percent. They rise and at 50 TeV signal losses of about 5\% have to be taken into account. The energy calibration is performed by means of through--going muons taking their energy deposition as standard. Electronic calibration is repeated in regular intervals of 6 months by injecting a calibration charge at the amplifier input. A stability of better than 2\% over two years of operation has been attained. The detector signal is shaped to a slow signal with $10~\mu$s risetime in order to reduce the amplifier noise to a level less than that of a passing muon. On the other hand, this makes a fast external trigger necessary. The principal trigger of KASCADE is formed by a coincident signal in at least five stations in one subgroup of 16 stations of the array. This sets the energy threshold to a few times $10^{14}$~eV depending on zenith angle and primary mass. An alternative trigger is generated by a layer of plastic scintillators positioned below the third iron layer at a depth of $2.2~\lambda_{I}$. These scintillators cover two thirds of the calorimeter surface and deliver timing information with 1.5~ns resolution. \section{Simulations} EAS simulations are performed using the CORSIKA versions 5.2 and 5.62 as described in \cite{corsika}. The interaction models chosen in the tests are VENUS version 4.12 \cite{venus}, QGSJET \cite{qgsjet} and SIBYLL version 1.6 \cite{sibyll}. We have chosen two models which are based on the Gribov Regge theory because their solid theoretical ground allows best to extrapolate from collider measurements to higher energies, forward kinematical regions, and nucleus--nucleus interactions. The DPMJET model, at the time of investigations, was not available in CORSIKA in a stable version. In addition, SIBYLL was used, a minijet model that is widely used in EAS calculations, especially as the hadronic interaction model in the MOCCA code. A sample of 2000 proton and iron--induced showers were simulated with SIBYLL and 7000 p and Fe events with QGSJET. With VENUS 2000 showers were generated, each for p, He, O, Si and Fe primaries. The showers were distributed in the energy range of 0.1\, PeV up to 31.6\, PeV according to a power law with a differential index of -2.7 and were equally spread in the interval of $15^o$ to $20^o$ zenith angle. In addition, the changing of the index to -3.1 at the {\it knee} position, which is assumed to be at 5~PeV, was taken into account. The shower axes were spread uniformly over the calorimeter surface extended by 2~m beyond its boundary. In order to determine the signals in the individual detectors, all secondary particles at ground level are passed through a detector simulation program using the GEANT package \cite{geant}. By these means, the instrumental response is taken into account and the simulated events are analysed in the same way as the experimental data, an important aspect to avoid systematic biases by pattern recognition and reconstruction algorithms. \section{Shower size determination} The data evaluation proceeds via three levels. In a first step the shower core and its direction of incidence are reconstructed and, using the single muon calibration of the array detectors, their energy deposits are converted into numbers of particles. In the next stage, iterative corrections for electromagnetic punch--through in the muon detectors and muonic energy deposits in the electron detectors are applied. The particle densities are fitted with a likelihood function to the Nishimura Kamata Greisen (NKG) formula \cite{kamata}. A radius parameter of 89~m and 420~m is used for electrons and muons, respectively. Because of limited statistics, the radial slope parameter (age) is fixed for the muons. The radius parameters deviate from the parameters originally proposed, but have been found to yield the best agreement with the data. The muon fit extends from 40~m to 200~m, the lower cut being imposed by the strong hadronic and electromagnetic punch--through near the shower centre. The upper boundary reflects the geometrical acceptance. In a final step, the muon fit function is used to correct the electron numbers and vice versa. The electromagnetic and muonic sizes $N_{e}$ and $N_{\mu}$ are obtained by integrating the final NKG fit functions. For the muons alternatively, integration within the range of the fit results in a truncated muon number $N'_\mu$. This observable has the advantage of being free of systematical errors caused by the extrapolation outside the experimental acceptance. As demonstrated in Fig.~\ref{eeichnmy}, it yields a good estimate of the primary energy irrespective of primary mass. To a certain extent, it is an integral variable indicating the sum of particles produced in the atmosphere independently of longitudinal cascade development. In the lefthand graph, the simulated values for the QGSJET model are plotted together with fitted straight lines. They show that in the $N'_\mu$ range given, the primary energy is proportional to the muon number $E_0 \propto N_\mu'^{0.98}$ with an error in the exponent of 0.06. This holds for the selected showers hitting the central detector with their axes. (For all showers falling into the area of the array a slightly higher coefficient of 1.10 is found.) It has been checked that the particle numbers are evaluated correctly up to values of $\lg N'_\mu=5$. At the highest energy of 100~PeV simulations indicate that $N'_\mu$ is overestimated by about 10\%. By studying $N_\mu$ sizes at this energy experimentally, irregularities in the muon size distribution may indicate an overestimation of 20\%. How well different models agree among each other is shown on the righthand part of Fig.~\ref{eeichnmy}, where the corresponding fitted lines are presented. It is seen that the SIBYLL model lies above the two others. In other words it generates fewer muons with consequences that will be discussed below. It is this truncated muon number $N'_\mu$ which we shall use throughout this article to classify events according to the muon number, that means approximately according to the primary energy. The accuracy of the reconstructed shower sizes is estimated to be 5\% for $N_{e}$ and 10\% for $N'_\mu$ around the {\it knee} position. \begin{figure}\centering \epsfig{file=eps/eeichnmy1_1.eps,width=0.48\textwidth}\hskip2mm% \epsfig{file=eps/eeichnmy2_1.eps,width=0.48\textwidth}% \caption{ Primary energy as determined by simulations from the truncated muon number $N'_\mu$ using the interaction models indicated. The vertical bars indicate the rms widths of primary energy distribution for fixed number of muons.} \label{eeichnmy} \end{figure} \section{Hadron reconstruction} The raw data of the central detector are passed through a pattern recognition program which traces a particle in the detector and reconstructs its position, energy and incident angle. Two algorithms exist. One of them is optimized to reconstruct unaccompanied hadrons and to determine their energy and angle with best resolution. The second is trained to resolve as many hadrons as possible in a shower core and to reconstruct their proper energies and angles of incidence. This algorithm has been used for the analyses presented in the following. Grosso modo, the pattern recognition proceeds as follows: Clusters of energy are searched to line up and to form a track, from which roughly an angle of incidence can be inferred. Then in the lower layers patterns of cascades are looked for since these penetrating and late developing cascades can be reconstructed most easily. Going upwards in the calorimeter, clusters are formed from the remaining energy and lined up to showers according to the direction already found. The uppermost layer is not used for hadron energy determination to evade hadron signals, which are too much distorted by the electromagnetic component, nor is the trigger layer used because of its limited dynamical range. Due to a fine lateral segmentation of 25~cm, the minimal distance to separate two equal--energy hadrons with 50\% probability amounts to 40~cm. This causes the reconstructed hadron density to flatten off at about 1.5~hadrons/m$^2$. The reconstruction efficiency with respect to the hadron energy is presented in Fig.~\ref{effi}. At 50~GeV an efficiency of 70\% is obtained. This energy is taken as threshold in most of the analyses in the following, if not mentioned otherwise. We present the values on a logarithmic scale in order to demonstrate how often high--energy radiating muons can mimic a hadron. Their reconstructed hadronic energy, however, is much lower, typically by a factor of 10. The fraction of non--identified hadrons above 100\,GeV typically amounts to 5\%. This value holds for a 1 PeV shower hitting the calorimeter at its centre and rises to 30\% at 10 PeV. This effect is taken into account automatically, because in the simulation it appears as the same token. \section{Event selection} About $10^8$ events were recorded from October 1996 to August 1998. In $6\times10^6$ events, at least one hadron was reconstructed. Events accepted for the present analysis have to fulfill the following requirements: More than two hadrons are reconstructed, the zenith angle of the shower is less than $30^{\circ}$ and the core, as determined by the array stations, hits the calorimeter or lies within 1.5 m distance outside its boundary. For shower sizes corresponding to energies of more than about 1 PeV, the core can also be determined in the first calorimeter layer by the electromagnetic punch--through. The fine sampling of the ionisation chambers yields 0.5~m spatial resolution for the core position. For events with such a precise core position it has to lie within the calorimeter at 1 m distance from its boundary. After all cuts, 40\,000 events were left for the final analysis. For non--centric showers, hadronic observables like the number of hadrons have been corrected for the missing calorimeter surface by requiring rotational symmetry. On the other hand, some variables are used, for which such a correction is not obvious, e.g. the {\it minimum--spanning--tree}, see Section 8.4. In these cases, only a square of $8\times8~\mbox{m}^2$ of the calorimeter with the shower core in its centre is used and the rest of the calorimeter information neglected. This treatment ensures that all events are analysed on the same footing. \section{Tests at large distances} \begin{figure}\centering \begin{minipage}[t]{0.5\textwidth} \epsfig{file=eps/effi.eps,width=\textwidth} \caption{Reconstruction efficiency for hadrons for two different zenith angles. The square symbols represent the probability of radiating muons misidentified as hadrons.} \label{effi} \end{minipage}% \begin{minipage}[t]{0.5\textwidth} \epsfig{file=eps/rho2.eps,width=\textwidth}% \caption{Density of hadrons (left scale) and of hadronic energy (right scale) for showers of truncated muon numbers as indicated corresponding to primary energies of 3--10~PeV. The curves represent fits of the NKG formula to the data.} \label{rho2} \end{minipage}% \end{figure} \begin{figure}\centering \epsfig{file=eps/rho_s1.eps,width=0.5\textwidth}% \epsfig{file=eps/rho_s2.eps,width=0.5\textwidth}% \caption{Density of hadronic energy (filled circles) vs. core distance for two intervals of primary energy. The indicated muon numbers correspond to $1~\mbox{PeV}\le E_0 < 3~\mbox{PeV}$ and $3~\mbox{PeV}\le E_0<10~\mbox{PeV}$. The CORSIKA simulations (open symbols) represent primary proton and iron nuclei with the QGSJET model.} \label{rho} \end{figure} \begin{figure} \epsfig{file=eps/ooty.eps,width=0.5\textwidth}% \epsfig{file=eps/tien.eps,width=0.5\textwidth} \caption{Lateral hadron density for electromagnetic shower sizes of $5.25 \le \lg N_{e} < 5.5$. Thresholds for hadron detection are 50~GeV (left), and 1~TeV (right). The dashed lines represent CORSIKA simulations with the QGSJET model for primary proton and iron nuclei using an exponential, see text.} \label{ootytien} \end{figure} \begin{figure}\centering \epsfig{file=eps/mlat1.eps,width=0.5\textwidth}% \epsfig{file=eps/mlat2.eps,width=0.5\textwidth}% \caption{The lateral hadron density for muonic shower sizes corresponding to a mean primary energy of 1.2 PeV. Threshold for hadron detection is 50~GeV. The data are compared with simulations using VENUS (left) and SIBYLL (right), the curves represent fits according to a modified exponential, see text.} \label{mlat} \end{figure} Studying hadron distributions at large core distances checks mainly the overall performance of the shower simulation program CORSIKA. In the regions far away from the shower axis of an EAS, the Monte Carlo calculations can be verified with respect to the transport of particles, their decay characteristics, etc. If the hadrons are well described it signifies that the shower propagation is treated properly. In these outer regions, where lower hadron energies and larger scattering angles dominate, the underlying physics is sufficiently well known from accelerator experiments, and the code in itself can be tested. As an example of such a test, the hadron lateral distribution is presented in Fig.~\ref{rho2} for $N'_\mu$ sizes corresponding to the primary energy interval around and above the {\it knee}: $3~\mbox{PeV}\le E_0<10~\mbox{PeV}$. The distributions of the number of hadrons and of the hadronic energy are given. In the very centre of the former, a saturation as mentioned in chapter 5 can be noticed. Several functions have been tried to fit the data points, among others exponentials as suggested by Kempa \cite{kempa}. However, by far the best fit was obtained when applying the NKG formula represented by the curves shown in the graph. This finding is not particularly surprising because hadrons of an energy of approximately 100~GeV \cite{kempa}, when passing through the atmosphere, generate the electromagnetic component, and the NKG formula has been derived for electromagnetic cascades. In addition, multiple scattering of electrons determining the Moli\`ere radius resembles the scattering character of hadrons with a mean transverse momentum of 400~MeV/c irrespective of their energy. Replacing the mean multiple scattering by the latter and the radiation length by the interaction length one arrives at a radius $R_H$ of about 10~m. This value we expect to take the place of the Moli\`ere radius in the NKG formula for electron measurements. Indeed, values of this order are found experimentally. Lateral hadron distributions compared with CORSIKA simulations are shown in Fig.~\ref{rho} for primary energies below and above the {\it knee}. In the diagrams the hadronic energy density is plotted for muon numbers corresponding to the primary energy intervals of $1~\mbox{PeV}\le E_0<3~\mbox{PeV}$ and $3~\mbox{PeV}\le E_0<10~\mbox{PeV}$. The data points are compared to primary proton and iron simulations applying the QGSJET model. These two extreme assumptions about the masses result in nearly identical hadron densities and the measured data coincide with the simulations, thereby verifying the calculations. Similar good agreement is found for the VENUS and SIBYLL models. Simulations and data agree well up to 100~m distance from the core. Only in the very inner region of 10~m, the simulations yield deviating hadron densities for different primary masses. Nevertheless, the measurements here lie well in between the two extreme primary compositions of pure protons or pure iron nuclei. \section{Tests at shower core} \subsection{Hadron lateral distribution} To begin with, the lateral distributions are compared with values published in the literature. Hadron distributions in the core of EAS have been measured at Ooty by Vatcha and Sreekantan \cite{vatcha} and at Tien Shan by Danilova et al. \cite{danilova}. Results of earlier experiments have been examined and discussed by Sreekantan et al. \cite{sreekantan}. In the experiments different techniques for hadron detection have been applied: a cloud chamber at Ooty, long gaseous ionisation tubes at Tien Shan and liquid ionisation chambers in the present experiment. Therefore, it is of interest to compare the respective results. The experiments were performed at different altitudes, and a priori they are expected to deliver deviating results. However, when compared at the same electromagnetic shower size, hadron distributions should be similar because electrons and hadrons, the latter of about 100 GeV, are closely related to each other in an EAS when the shower passes through the atmosphere. A sort of equilibrium turns up as has been pointed out by Kempa \cite{kempa}. Indeed, Fig.~\ref{ootytien} demonstrates for electron numbers $5.25 \leq \lg N_{e} < 5.5$ that the lateral hadron distributions agree reasonably well. In particular, the measurements of the Ooty group at an atmospheric depth of $800~\mbox{g}/\mbox{cm}^2$ coincide with the present findings. The grey shaded band represents CORSIKA simulations using the hadronic interaction model QGSJET, the lower curve representing primary protons and the upper curve primary iron nuclei. The curves are fits to the simulated density of hadrons according to $\rho_H(r)\propto exp~(-(\frac{r}{r_0})^\kappa)$ with values for $\kappa$ found to be between 0.7 and 0.9. The data lie well between these two boundaries. The graph on the righthand side represents hadron densities with a threshold of 1 TeV. Bearing this high threshold in mind, the similarity in both distributions, Tien Shan at $690~\mbox{g}/\mbox{cm}^2$ and KASCADE at sea--level, is astonishing. In conclusion, it can be stated that hadron densities, despite of being measured with different techniques, agree reasonably well among different experiments. When classifying hadron distributions according to muonic shower sizes, differences among the interaction models emerge. This becomes apparent in Fig.~\ref{mlat}, where the central density is plotted for truncated muon numbers which correspond to a mean energy of about 1.2 PeV. On the left graph, the VENUS calculations enclose the data points leaving the elemental composition to be somewhere between pure proton or pure iron primaries. On the right graph, the measured data points follow the lower boundary of the SIBYLL calculations, suggesting that all primaries are iron nuclei, at this energy obviously an unprobable result. The lateral distribution demonstrates, and other observables in a similar manner as reported previously \cite{jrh}, that the SIBYLL code generates too low muon numbers thereby entailing a comparison at a different estimate of the primary energy. A hint has already been observed in Fig.~\ref{eeichnmy} where the SIBYLL lines lie above those of QGSJET and VENUS. When hadronic observables are classified according to electromagnetic shower sizes, the disagreement vanishes as will be discussed in the following. \subsection{Hadron energy distribution} \begin{figure}\centering \begin{minipage}[t]{0.5\textwidth}% \epsfig{file=eps/mespec4.eps,width=\textwidth}% \caption{Hadron energy distribution for fixed electron number $N_{e}$ corresponding approximately to 6 PeV primary energy. The lines represent CORSIKA simulations with three interaction models, the lower curves for primary iron, the upper for protons.} \label{mespec} \end{minipage}% \begin{minipage}[t]{0.5\textwidth}% \epsfig{file=eps/nintne2.eps,width=\textwidth}% \caption{Hadronic shower size vs. electromagnetic shower size. Experimental values are compared with simulations using VENUS (full lines) and QGSJET (shaded area), both for primary protons and iron nuclei. The experimental error bars are rms--values.} \label{nintne} \end{minipage}% \end{figure} The energy distribution of hadrons is shown in Fig.~\ref{mespec} for a fixed electromagnetic shower size. Plotted is the number of hadrons in an area of $8\times8~\mbox{m}^2$ around the shower core. As already mentioned, in this way all showers are treated in the same manner, independent of their point of incidence. To avoid a systematic bias the loss in statistics has to be accepted. The number of showers reduces to about 5000. The shower size bin of $5.5 \leq \lg N_{e} < 5.75$ corresponds approximately to a mean primary energy of 6~PeV. The lines represent fits to the simulations according to $exp~(-(\frac{\lg E_H-a}{b})^c)$. Usually in the literature $c=1$ is assumed, however, the present data, due to their large dynamical range, yield values for $c$ from 1.3 to 1.6. As can be inferred from the graph, all three interaction models reproduce the measured data reasonably well, elucidating the fact that electrons closely follow the hadrons in EAS propagation. But if the same data are classified corresponding to the muon number, again SIBYLL seems to generate too many hadrons and thereby mimics a primary composition of pure iron nuclei. For this reason SIBYLL will not be utilized any further. In the figure also the energy spectrum is plotted as measured with the Maket--ANI calorimeter by Ter--Antonian et al.\ \cite{ter}. As already mentioned above, distributions are expected to coincide when taken at the same electron number even if they have been measured at different altitudes. In the present case the data have been taken at sea level and at $700~\mbox{g}/\mbox{cm}^2$ on Mount Aragats. The energy distributions, indeed, agree rather well with each other indicating that in both data sets the patterns of hadrons are well recognized and the energies correctly determined. It was seen that SIBYLL encounters difficulties when the data are classified according to muonic shower sizes. The model VENUS, on the other hand, cannot reproduce hadronic observables convincingly well when they are binned into electron number intervals. An example is given in Fig.~\ref{nintne}. It shows the number of hadrons, i.e. the hadronic shower size $N_h$, as a function of the electromagnetic shower size $N_e$. The experimental points match well to the primary proton line as expected from QGSJET predictions. This phenomenon is easily understood by the steeply falling flux spectrum and the fact that primary protons induce larger electromagnetic sizes at observation level than heavy primaries. Hence, when grouping in $N_{e}$ bins , showers from primary protons will be enriched and we expect to have predominantly proton showers in our sample. This fact reduces any ambiguities in the results due to the absence of direct information on primary composition. Concerning the VENUS model the predicted hadron numbers are too high and the two lines which mark the region between primary protons and iron nuclei cannot explain the data. The point at the lowest shower size is still influenced by the trigger efficiency of the array counters. \subsection{Hadron energy fraction} \begin{figure}\centering \epsfig{file=eps/mefracl1.eps,width=0.5\textwidth}% \epsfig{file=eps/mefracl2.eps,width=0.5\textwidth}% \caption{The energy fraction of all hadrons vs. the most energetic hadron in a shower. The data are compared to simulations using the QGSJET model for primary protons (p) and iron nuclei (Fe). Shaded is the physically meaningful region as obtained from the simulations. Primary energies correspond to 2 PeV (left) and 12 PeV (right).} \label{mefracl} \end{figure} A suitable test of the interaction models consists in investigating the granular structure of the hadronic core concerning spatial as well as energy distributions. As variables we have chosen the energy fraction of hadrons and the distances in the {\it minimum--spanning--tree} between them. Both will be dealt with in the following sections. For each hadron its energy fraction with respect to the most energetic hadron in that particular shower is calculated. For primary protons, the leading particle effect is expected to produce one particularly energetic hadron accompanied by hadrons with a broad distribution of lower energies. Hence, we presume to find a rather large dispersion of hadronic energies for primary protons, whereas for primary iron nuclei the hadron energies should be more equally distributed. The simulated distributions, indeed, confirm this expectation as is shown in Fig.~\ref{mefracl}. The lines --- to guide the eye --- represent fits to the simulations using two modified exponentials as in the preceding section, which are connected to each other at the maximum. On the lefthand graph, the data seem to corroborate the simulations. They are shown for a muon number range corresponding to a primary energy of approximately 2 PeV, i.e., below the {\it knee} position. On the righthand side, the results are shown for an interval above the {\it knee} for muonic shower sizes corresponding to a primary energy of 12 PeV. The reader observes that the data cannot be explained by the simulations, neither by primary protons nor by iron nuclei. On a logarithmic scale the data exhibit a symmetric distribution around the value $\lg (E_H/E_H^{max})\cong-1.5$, even more symmetric than would be expected for a pure iron composition. In particular, energetic hadrons resulting from the leading particle effect seem to be missing. They would shift the distribution to smaller values. This absence of energetic hadrons in the observations will be confirmed later when investigating other observables. \subsection{Minimum--spanning--tree} \begin{figure}\centering \epsfig{file=eps/mst35.eps,width=0.5\textwidth}% \caption{Example of a hadronic core in the calorimeter (top view). The square marks the acceptance area of $8\times8~\mbox{m}^2$ around the shower centre (star). The energy of each hadron is indicated by the area of its point in a logarithmic scale.} \label{mst} \end{figure} \begin{figure}\centering \epsfig{file=eps/mmsts1.eps,width=0.5\textwidth}% \epsfig{file=eps/mmsts2.eps,width=0.5\textwidth}% \caption{The distances in a minimum--spanning--tree for a muon number interval below (left) and above (right) the {\it knee}. The measurements are compared with simulations using the interaction model QGSJET for primary protons (p) and iron nuclei (Fe). The lines are fits to the simulations analogous to Fig.~\ref{mefracl}.} \label{mmsts} \end{figure} When constructing the {\it minimum--spanning--tree} (MST), all hadrons are connected to each other in a plane perpendicular to the shower axis. The MST is that configuration where the sum of all connections weighted by the inverse energy sum of its neighbours has a minimum. The $1/E$ weighting has been found to separate iron and proton induced showers the most. Fig.~\ref{mst} shows as an example the central shower core of an event. Plotted are the points of incidence on the calorimeter. The sizes of the points mark the hadron energies on a logarithmic scale. The shower centre and the fiducial area of $8\times8~\mbox{m}^2$ around it are indicated as well. For each event the distribution of distances is formed. Average distributions from many events are given in Fig.~\ref{mmsts}. As in Fig.~\ref{mefracl}, the muonic shower sizes correspond to primary energy intervals below and above the {\it knee}. It is observed that for the former, the data lie well within the bounds of the primary composition but that above the {\it knee} the measurements yield results which are not in complete agreement with the model although they are close to the simulated iron data. The distributions of Figs.~\ref{mefracl} and \ref{mmsts} have also been calculated analysing the full calorimeter surface and not only the $8 \times 8~\mbox{m}^2$ around the shower centre. No remarkable difference could be obtained. In both observables -- energy fraction and MST -- the data for higher primary energies cannot be interpreted by the simulations. Additionally, in the M.C. calculations the {\it knee} in the primary energy distribution has been omitted. Again, no remarkable change in the distributions showed up. In fact, when investigating the distributions as a function of muon number, the deviation between M.C. values and the measured data develops smoothly with increasing energy. When regarding the righthand graph in Fig.~\ref{mmsts} the question arises whether the interaction model produces too small distances or too energetic hadrons or both. In agreement with the observation in Fig.~\ref{mefracl} one has to conclude that too energetic hadrons are generated as compared to the data. Whether in the MSTs also the distances between the hadrons, in other words the transverse momenta, are underestimated cannot be decided at the moment. Also the number of hadrons plays a role. This issue is under further investigations. \subsection{Hadronic energy in large showers} \begin{figure}\centering \epsfig{file=eps/nintnmy2_2.eps,width=0.5\textwidth}% \epsfig{file=eps/me1nmy_3.eps,width=0.5\textwidth} \caption{Hadronic shower size vs. muonic shower size (left) and the maximum hadron energy in a shower vs. its muonic shower size (right). The lines represent simulations with the indicated interaction codes. The upper curves represent primary protons, the lower primary iron nuclei.} \label{nintnmy} \end{figure} \begin{figure}\centering \epsfig{file=eps/hzahl1.eps,width=0.5\textwidth}% \caption{Hadron number at different thresholds, 50 GeV (diamonds), 200 GeV (triangles), and 1 TeV (squares) vs. electromagnetic shower size. The dashed band represents QGSJET simulations for primary protons (lower line) and iron nuclei (upper line).} \label{hzahl} \end{figure} Deviations between measurement and simulations as in the preceding sections are also observed when investigating the hadronic energy in large showers. With rising muon numbers $N'_\mu$, the experiment reveals an increasing part of missing hadronic energy in the shower core. Fig.~\ref{nintnmy} (left) shows the number of hadrons versus the muonic shower size. At muon numbers corresponding to about 5 PeV primary energy, the hadron numbers turn out to be smaller than predicted for iron by both interaction models VENUS and QGSJET. Again, one observes that the latter model describes the experimental points somewhat better. The conclusion that QGSJET reproduces the data best in the PeV region is also confirmed by a recent model comparison performed by Erlykin and Wolfendale \cite{wolfendale}. The authors classify the models on the basis of consistency checks among different observables, e.g.\ the depth of shower maximum $X_{max}$ and the $N_\mu / N_e$ ratio. The righthand graph of Fig.~\ref{nintnmy} presents the maximum hadron energy found in showers with the indicated muon number. The open symbols represent the QGSJET simulations, again QGSJET and VENUS yield similar results. Measurement and simulation disagree also to some extent in this variable at large shower sizes. The overestimation of muon numbers mentioned in section 4 cannot account for the discrepancies. On a logarithmic scale it starts to be noticeable at $\lg N'_\mu=5.5$ and amounts to $\Delta\lg N'_\mu=0.1$. A shift of this size does not ameliorate the situation. The data have been checked independently in the reduced fiducial area of $8\times8~\mbox{m}^2$. But in this analysis, too, the data seem to fake pure iron primaries at $\lg N'_\mu=4.3$ and are below that boundary for larger muonic sizes. Factum is that we do not observe the energetic hadrons expected from the M.C. calculations. In the energy region 10 to 100~PeV even QGSJET fails to describe the measurements. Obviously, the question arises whether these experimentally detected effects are artifacts caused, for instance, by saturation effects in the calorimeter or by insufficient pattern recognition presumed to be different in the simulation than present in the experimental data. After all, the high--energy values correspond to primary energies of about 100 PeV where 400 hadrons have to be reconstructed. At this point, it may be noted again that always the experimental and simulated data are compared with each other on the detector signal level, hence, a possible hadron misidentification applies to both data sets. As already pointed out in section 2, individual hadrons up to 50~TeV have been reconstructed and their saturation effects have been examined thoroughly. Some misallocation of energy to individual hadrons might occur, though, if lateral distributions of hadrons in the core differ markedly between simulations and reality. There may be indication from emulsion experiments for this \cite{tamada}. However, from the results shown in Fig.~\ref{ootytien} and \ref{mlat} we would not expect any dramatic effect. Fig.~\ref{hzahl} demonstrates that for large electromagnetic shower sizes, the number of hadrons compares well with other experiments as well as with CORSIKA simulations. In the diagram the number of hadrons above the indicated thresholds is presented with respect to the shower size. The values obtained for hadrons above 1~TeV can be related to two other experiments performed at Kiel by Fritze et al. \cite{fritze} and on the Chacaltaya by C. Aguirre et al. \cite{aguierre}. It is observed that up to shower sizes which correspond to about 20 PeV for primary protons all high--energy hadrons are reconstructed, i.e., more than 70~TeV energy are found in the calorimeter. When compared to QGSJET simulations, the data lie within the physical boundaries as shown for the 1~TeV line. On closer inspection the data indicate an increase of the mean mass with rising energy. Also in Fig. \ref{nintne} it has been seen that the hadron numbers are well reproduced by QGSJET up to the highest electromagnetic shower sizes. In conclusion, it can be stated that the hadron component compares well between different experiments and with M.C. calculations when classified according to electromagnetic shower sizes, and that the deviations observed in muon number binning cannot be accounted for by experimental imperfections. \section{Conclusion and outlook} Three interaction models have been tested by examining the hadronic cores of large EAS. It turned out that QGSJET reproduces the data best, but at large muonic shower sizes, i.e.\ at energies above the {\it knee} even this model fails to reproduce certain observables. Most importantly, the model predicts more hadrons than are observed experimentally. The current investigation is a first approach with a first data sample of the KASCADE experiment. Better statistics, both in the data and in the Monte Carlo calculations, are imperative, especially above the {\it knee} in the 10~PeV region, and are expected from the further operation of the experiment. In addition, other experimental methods have to be developed to check the simulation codes even more rigorously. Such a stringent check consists of verifying absolute particle fluxes at ground level at energies where the primary flux is reasonably well known. Improvements in the interaction models are also under way. {\small NE}X{\small US} is in statu nascendi, a joint enterprise by the authors of VENUS and QGSJET \cite{werner}. It has become evident that a very precise description of the shower development in the atmosphere is needed if the mass of the primaries is to be estimated by means of ground level particle distributions. \section{Acknowledgments} The authors would like to thank the members of the engineering and technical staff of the KASCADE collaboration who contributed with enthusiasm and engagement to the success of the experiment. The Polish group gratefully acknowledges support by the Polish State Committee for Scientific Research (grant No. 2 P03B 16012). The work has been partly supported by a grant of the Rumanian Ministry of Research and Technology and by the research grant no. 94964 of the Armenian Government and ISTC project A 116. The support of the experiment by the Ministry for Research of the German Federal Government is gratefully acknowledged. \section*{References}
\section{Introduction} Linearization of a dynamical system near a periodic orbit is one of the most fruitful starting points for the analysis of classical motion. In its turn, the symplectic group of linear Hamiltonian systems in plane phase space is easily quantized to form the corresponding metaplectic quantum group. Essentially, the generating function for the group of canonical transformations is simply exponentiated to obtain a representation of the quantum unitary transformation. If the chosen orbit is a point of equilibrium, the corresponding linear system belongs to the homogeneous symplectic group, characterized by a single equilibrium, usually taken as the origin. Likewise, the Poincar\'{e} map in the neighborhood of a periodic orbit is linearized into a homogeneous symplectic map with discrete time. The essential character of the motion is classified according to the eigenvalues of the symplectic matrix ${\cal M}$, that determines the evolution of phase space points $x:$% \begin{equation} x^{\prime }={\cal M}x. \end{equation} There may be \begin{description} \item[(a)] pairs of eigenvalues ($\lambda ,\lambda ^{-1});$ \item[(b)] pairs of eigenvalues on the unit circle ($e^{i\theta },e^{-i\theta })$ \item[(c)] quartets of general complex eigenvalues $\lambda ^{\pm 1}e^{\pm i\theta }.$ \end{description} \noindent On varying parameters, it is possible to obtain unit eigenvalues , or eigenvalue collisions, but the above classification is generic for a given symplectic system \cite{quinze}. It is always possible to decompose such a generic linearized system into sub-systems in invariant subspaces of two dimensions, corresponding to cases (a)\ and (b) above, or four dimensions in case (c). Case (b) is the {\it % elliptic map, }which is trivially integrable, whereas case (a)\ defines {\it % hyperbolic motion.} This is also very simple in the linear limit, but can become a source of chaotic mixing as nonlinear perturbations are added. Alternatively, this effect is achieved by wrapping the plane space itself into a torus. The resulting symplectomorphism of the torus is known colloquially as a {\it % cat map}, characterized by a symplectic matrix with integer elements. A hyperbolic cat map is structurally stable, i.e. the orbit structure is invariant with respect to small nonlinear perturbations as a consequence of Anosov's theorem \cite{quinze}. The same is true of a four dimensional {\it % loxodromic cat map} with general complex eigenvalues in case (c). These structurally invariant systems are known as {\it Anosov} systems, they are ergodic and mixing. It follows that four dimensions is the lower bound in which we can study loxodromic periodic orbits \cite{quinze}, characterized by stable and unstable manifolds where the orbits spiral inwards and outwards respectively, and their effect on the quantum energy spectrum. This is the reason for their absence in all previous studies of the quantization of cat maps, though Greenman \cite{greenman} has recently analyzed the periodic structure of higher dimensional classical cat maps. Dimension four is also the least dimension for the analysis of the decomposition of the neighborhood of orbits into elliptic and (real)\ hyperbolic components. In some cases this decomposition is only local, because the canonical transformation that achieves it is not itself a cat map. Then the quantum quasienergy spectrum will not be decomposed into the corresponding lower dimensional spectra. In any case, all cat maps derived from each other as a result of similarity transformations involving other cat maps are equivalent: they have the same (classical and quantum) eigenvalues and the same number of fixed points. ( Note that, on the torus, a homogeneous linear map may have multiple fixed points.) For this reason, we discuss in the next section the integer subgroup of symplectic transformations, nicknamed the feline group. For higher dimensions than two, we encounter the problem of a priori identification of a cat map. An alternative approach involves generating functions, of which there are several choices. However there is a great advantage to using generating functions that are invariant with respect to feline transformations. In section 3 we analyze the dynamics of classical cat maps and classify four-dimensional cat maps, providing examples of various types. These examples are then quantized in section 4. They share a simplifying property that permits us to discuss the periodicity of the propagator and the exactness of the Gutzwiller trace formula, without analyzing the subtleties of general cat map quantization. This is the subject of section 5, where we determine the set of Floquet angles that allow the quantization of a given map; and study the feline invariance of the quantization. \section{ Feline-invariant generating functions} \setcounter{equation}{0} A point in the even-dimensional {\it phase space} with $L$ degrees of freedom on a $2L$-torus has coordinates separated into $L$ momenta and $L$ positions, so that $x=\left( \begin{array}{c} p \\ q \end{array} \right) =\left( \begin{array}{c} p_1,\cdots ,p_L \\ q_1,\cdots ,q_L \end{array} \right) $. All the $2L$ coordinates are periodic with periods $\Delta q_i$ and $\Delta p_i$. For simplicity we will treat the case where we can choose units so that $\Delta q_i$ and $\Delta p_i$ are all equal to 1. The range of values of $x$ is then the unit $2L$-hypercube denoted from now on by $\Box .$ Let us consider then a linear automorphism on the $2L$-torus generated by the $2L\times 2L$ matrix ${\cal M}$, that takes a point $x_{-}=\left( \begin{array}{c} p_{-} \\ q_{-} \end{array} \right) $ to a point $x_{+}=\left( \begin{array}{c} p_{+} \\ q_{+} \end{array} \right) :$% \begin{equation} x_{+}={\cal M}x_{-}\quad \mbox{mod(1)}. \label{mapa0} \end{equation} In other words, there exists an integer $2L$-dimensional vector ${\bf m}% =\left( \begin{array}{l} m_p \\ m_q \end{array} \right) ,$ such that \begin{equation} x_{+}={\cal M}x_{-}-{\bf m.} \label{mapa} \end{equation} The components of ${\bf m}$ denote the winding numbers made by the point $% x_{-}$ around the respective irreducible circuit on the $2L$-torus after the application of the map ${\cal M}.$ The torus will be divided into regions labeled by their respective vector ${\bf m}$. For the map to be conservative, the ${\cal M}$ matrix must be symplectic, that is \begin{equation} {\cal M}^t{\frak J}{\cal M}={\frak J,} \label{msimple} \end{equation} where ${\cal M}^t$ is the transpose of ${\cal M}$ and \begin{equation} {\frak J}=\left[ \begin{array}{c|c} 0 & -1 \\ \hline 1 & 0 \end{array} \right] . \end{equation} The matrix ${\cal M}$ must have integer coefficients for the $2L$-torus to be mapped onto itself. For one degree of freedom ( $L=1$ ) these systems are known as Arnold's cat maps \cite{quinze}. If $|tr({\cal M})|>2$ the map has two distinct real eigenvectors, so it is ergodic, mixing and purely hyperbolic. For the case $% |tr({\cal M})|<2$ there are no real eigenvectors, the map is then elliptic. For $% |tr({\cal M})|=2$ there is only one real eigenvector and so the map is parabolic. For more degrees of freedom richer structure may appear, as will be discussed. The set of matrices ${\cal M}$ satisfying (\ref{msimple}) form the symplectic group, so that the matrix ${\cal M}^{\prime },$ obtained from ${\cal M}$ by the similarity transformation \begin{equation} {\cal N}^{-1}{\cal M}{\cal N}={\cal M}^{\prime } \label{2.5} \end{equation} is also symplectic if the matrix ${\cal N}$ is itself symplectic and shares the same eigenvalues as ${\cal M}$. Indeed, we may consider the symplectic transformation corresponding to ${\cal M}^{\prime }$ as the same as ${\cal M}$, but viewed in an alternative symplectic coordinate system. Consider now a product of cat maps; this must be symplectic and all the matrix elements will be integers. Since the inverse of a cat map is also a symplectic integer transformation and the unit matrix likewise, the set of cat maps is a subgroup of the symplectic group, appropriately nicknamed the {\it feline group}. Again, we may consider similarity transformations of the form (\ref{2.5}) among cat maps ${\cal M}$ and ${\cal N}$ as defining essentially maps viewed by alternative symplectic coordinates on the torus. The importance of defining equivalence classes of similar cat maps increases with the dimension of the torus. It is easy to define integer matrices and the {\it a posteriori} condition (\ref{msimple}) merely restricts the value of the determinant to unity when $L=1.$\ However, for $L=2$, the symplectic property already implies ten independent conditions to be satisfied by the sixteen integer matrix elements. An alternative procedure, that we adopt here, is to define the transformation implicitly by means of a generating function, thus guaranteeing (\ref{msimple}). The disadvantage is now that we need to ensure that the resulting matrix ${\cal M}$ has integer entries, leading to additional restrictions on the allowed generating functions (see below ) . Nevertheless, this approach has proved very fruitful, because the quantization relies heavily on the generating function, and not on the transformation matrix itself. The generating function in position representation for the one degree of freedom case is \cite{keat1} \begin{equation} S(q_{-},q_{+},{\bf m})=\frac 1{2{\cal M}_{21}}\left[ {\cal M}_{22}q_{-}^2-2q_{-}(q_{+}+m_q)+{\cal M}_{11}(q_{+}+m_q)^2-2{\cal M}_{21}m_pq_{+}\right] . \end{equation} This function generates the dynamics through \begin{eqnarray} p_{+} &=&\frac{\partial S(q_{-},q_{+},{\bf m})}{\partial q_{+}} \\ p_{-} &=&-\frac{\partial S(q_{-},q_{+},{\bf m})}{\partial q_{-}}. \end{eqnarray} Thus the quadratic part of the generating function is common to the entire torus, whereas the linear part depends on the winding vector ${\bf m}$ that changes discontinuously on the boundary of each subregion of the torus. The weakness of the position generating function is that it transforms in a complicated way under feline similarity transformations. Only in the special case of a point transformation, that does not mix momenta with positions, will the generating function remain invariant. In contrast, the {\it center and chord generating functions}, are known to be symplectically invariant \cite{ozrep}. Adapted to the torus, we shall show that they can also be chosen to be invariant under feline transformations. The starting point is to define the center point $x$ and the chord $\xi $ such that \begin{equation} x_{\pm }=x\pm \frac 12\xi , \label{qqxcor} \end{equation} so that the center \begin{equation} x\equiv \frac{x_{+}+x_{-}}2 \label{xdef} \end{equation} and the chord \begin{equation} \xi \equiv x_{+}-x_{-}. \label{cordef} \end{equation} Given the initial point $x_{-}$, \begin{eqnarray} x &=&\frac 12\left( {\cal M}+1\right) x_{-}-\frac{{\bf m}}2 \label{xx-} \\ \xi &=&\left( {\cal M}-1\right) x_{-}-{\bf m.} \label{corx-} \end{eqnarray} Elimination of $x_{-}$, then establishes the direct relation between centers and chords: \begin{eqnarray} x &=&\frac 12\frac{\left( {\cal M}+1\right) }{({\cal M}-1)}\xi +({\cal M}-1)^{-1}{\bf m} \label{xcor} \\ \xi &=&2\frac{\left( {\cal M}-1\right) }{\left( {\cal M}+1\right) }x-2\left( {\cal M}+1\right) ^{-1}{\bf m}. \label{corx} \end{eqnarray} Center and chord generating functions respectively denoted by $S(x,{\bf m})$ and $S(\xi ,{\bf m})$ are defined in \cite{ozrep} so that the transformation is obtained as \begin{eqnarray} x &=&{\frak J}\frac{\partial S(\xi ,{\bf m})}{\partial \xi } \label{xgenfu} \\ \xi &=&-{\frak J}\frac{\partial S(x,{\bf m})}{\partial x}. \label{corgenfu} \end{eqnarray} In analogy to the dynamics in the plane, we may consider that use of the center representation identifies the orbit with the reflection (or inversion) through $x$, because of (\ref{xdef}). Hence we shall also refer to $x$ as the reflection point. The chord representation is locally equivalent to the uniform translation of phase space by the chord (\ref {cordef}). Equating (\ref{xcor}) with (\ref{xgenfu}) and (\ref{corx}) with (% \ref{corgenfu}), we obtain the quadratic generating functions \begin{eqnarray} S(x,{\bf m}) &=&xBx+x(B-{\frak J)}{\bf m}+f({\bf m}) \label{sxf} \\ S(\xi ,{\bf m}) &=&\frac 14\xi \beta \xi +\frac 12\xi (\beta +{\frak J)}{\bf % m}+g({\bf m}) \label{scorg} \end{eqnarray} where $B$ and $\beta $ are symmetric matrices; they are the Cayley parametrization of ${\cal M}:$% \begin{eqnarray} {\frak J}B &=&\frac{\left( 1-{\cal M}\right) }{\left( 1+{\cal M}\right) } \label{jbm} \\ {\frak J}\beta &=&\frac{\left( {\cal M}+1\right) }{\left( {\cal M}-1\right) } \label{jbetam} \end{eqnarray} If ${\cal M}$ has an eigenvalue equal to $1$ then the $\beta $ matrix will be singular. This corresponds to a caustic of the center generating function. Whereas if ${\cal M}$ has an eigenvalue equal to $-1$, then $B$ will be a singular matrix which corresponds to a caustic of the chord generating function \cite {ozrep}.Some useful relations obtained from (\ref{jbm}) and (\ref{jbetam}) are \begin{eqnarray} (B-{\frak J}) &=&-2{\frak J}\left( 1+{\cal M}\right) ^{-1}, \\ (\beta +{\frak J}) &=&-2{\frak J}\left( {\cal M}-1\right) ^{-1}, \end{eqnarray} \begin{equation} {\frak J}B=-\frac 1{{\frak J}\beta } \label{jbbeta} \end{equation} and \begin{equation} {\cal M}=\frac{\left( 1-{\frak J}B\right) }{\left( 1+{\frak J}B\right) }=\frac{% \left( {\frak J}\beta +1\right) }{\left( {\frak J}\beta -1\right) }. \label{mbbeta} \end{equation} The functions $f({\bf m})$ and $g({\bf m})$ are arbitrary, since they only depend on the winding number ${\bf m}$, so they do not affect the transformation (\ref{xcor}) or (\ref{corx}). However, the center generating function for cat maps can also be obtained directly from the map (\ref{mapa}% ). This is a composition of the symplectic map on the plane ${\cal M}$ , whose center generating function is $S_1(x)=xBx,$ with the uniform translation $% T_{-{\bf m}}$ of vector $-{\bf m}$ that pulls back the final point to the unit cell $\Box .$ The generating function of such a translation is \cite {ozrep} $S_2(x)=-{\bf m\wedge }x$ , also symplectically invariant. Then, using the composition law for center generating functions \cite{ozrep} we obtain the generating function (\ref{sxf}) with \begin{equation} f({\bf m})=\frac 14{\bf m}B{\bf m.} \label{3.30} \end{equation} As in the plane case \cite{ozrep} generating functions are related among themselves by Legendre transformations. Thus, $S(x,{\bf m})$ is obtained from the more familiar position generating function $S(q_{-},q_{+},{\bf m})$ as \begin{equation} S(x,{\bf m})=S(q_{-},q_{+},{\bf m})+\frac 12(p_{-}+p_{+})(q_{+}-q_{-}), \label{legqx} \end{equation} whereas the relation between the chord and center generating function is \begin{equation} S(\xi ,{\bf m})=\xi \wedge x-S(x,{\bf m}). \label{legxcor} \end{equation} In each case, the variable absent on the left side is eliminated by requiring the right side to be stationary with respect to it. The skew product in (\ref{legxcor}) , \begin{equation} \xi \wedge \eta \ \equiv \sum_{\ell =1}^L\left( \xi _{p_\ell }\eta _{q_\ell }-\xi _{q_\ell }\eta _{p_\ell }\right) =({\frak J}\xi ).\eta =-\xi {\frak J}% \eta , \end{equation} is the symplectic area of the parallelogram formed by any pair of vectors $% \xi $ and $\eta $. Then, using (\ref{legxcor}), for the center function with the term (\ref{3.30}), we obtain the chord generating function (\ref{scorg}) with \begin{equation} g({\bf m})=\frac 14{\bf m}\beta {\bf m.} \label{3.30g} \end{equation} We shall label periodic points of period $l$ as $x_l.$ Thus, fixed points $% x_1$ are such that the chord $\xi =0,$ or the center $x=x_1=({\cal M}-1)^{-1}{\bf m,% }$ which inserted in (\ref{legxcor}) leads to \begin{equation} S(\xi =0,{\bf m})=-S(x_1,{\bf m}). \label{spcorx} \end{equation} There follows the restriction that the terms in $S(\xi ,{\bf m})$ and $S(x,% {\bf m})$ that depend only on ${\bf m}$ satisfy \begin{equation} f({\bf m})+g({\bf m})=\frac 14{\bf m}(\beta +B){\bf m.} \end{equation} The choice (\ref{3.30}) and (\ref{3.30g}) obviously satisfy this criterion, but another possibility is \begin{equation} f({\bf m})=\frac 14{\bf m}(B+\widetilde{{\frak J}}){\bf m\quad }\mbox{and}% \quad g({\bf m})=\frac 14{\bf m}(\beta -\widetilde{{\frak J}}){\bf m,} \label{gm} \end{equation} where we define the symmetric matrix \begin{equation} \widetilde{{\frak J}}=\left[ \begin{array}{c|c} 0 & 1 \\ \hline 1 & 0 \end{array} \right] . \end{equation} Using (\ref{gm}), we match the value of the action for a fixed point previously proposed by Keating \cite{keat1}, for the position representation \begin{equation} S(q_{-}=q_f,q_{+}=q_f,{\bf m})=S(x_1,{\bf m}). \end{equation} In conclusion, the center and chord generating functions for multidimensional cat maps are \begin{eqnarray} S(x,{\bf m}) &=&xBx+x(B-{\frak J)}{\bf m}+\frac 14{\bf m}(B+\widetilde{% {\frak J}}){\bf m} \label{sx} \\ S(\xi ,{\bf m}) &=&\frac 14\xi \beta \xi +\frac 12\xi (\beta +{\frak J)}{\bf % m}+\frac 14{\bf m}(\beta -\widetilde{{\frak J}}){\bf m.} \label{scor} \end{eqnarray} The corresponding generating functions for the transformation $x_{+}={\cal M}x_{-}$ in the plane are just $S(x,0)$ and $S(\xi ,0)$. It is important to note that all the reflections of the torus can be obtained with center points whose coordinates are in $[0,\frac 12].$ It would thus be possible to define such center points $x^{\prime }$ as \begin{equation} x^{\prime }\equiv \frac{x_{+}+x_{-}}2\qquad \mbox{mod}(\frac 12). \end{equation} This choice would not lead to the explicit relation (\ref{xx-}) with the winding number of the transformation, but instead \begin{equation} x^{\prime }=\frac 12\left( {\cal M}+1\right) x_{-}-\frac{{\bf m}^{\prime }}2, \label{xpm} \end{equation} where ${\bf m}^{\prime }$ has coordinates ${\bf m}_i^{\prime }=({\bf m}_i$ or ${\bf m}_i{\bf +1}),$ so that all the coordinates of $x^{\prime }$ would be in $[0,\frac 12].$ Hence, we allow instead the center points $x,$ defined as in (\ref{xdef}), to have coordinates in the full interval $[0,1],$ keeping the explicit relation with the winding number of the transformation. As a consequence, the chords defined as in (\ref{cordef}),\ have coordinates lying in the extended range $[-1,1].$ Thus, center points differing by integer loops around the torus are equivalent and so are chords differing by two integer loops. \begin{eqnarray} x &\equiv &x+{\bf k} \label{xequi} \\ \xi &\equiv &\xi +2{\bf k.} \label{corequi} \end{eqnarray} We find that (\ref{xequi}) and (\ref{corequi}) imply that, in (\ref{xcor}) and (\ref{corx}) respectively, the winding number ${\bf m}\,$ is equivalent to: \begin{equation} {\bf m}\equiv {\bf m}^{\prime }={\bf m}+({\cal M}-1){\bf k} \label{mprim} \end{equation} in (\ref{xcor}) and \begin{equation} {\bf m}\equiv {\bf m}^{\prime \prime }={\bf m}-({\cal M}+1){\bf k}. \label{m2prim} \end{equation} in (\ref{corx}). This implies that replacing $\,$ ${\bf m}$ by ${\bf m}% ^{\prime }$ in the generating function $S(\xi ,{\bf m})$ will lead to equivalent center points related by (\ref{xequi}). To obtain equivalent chords related by (\ref{corequi}), it is necessary to replace ${\bf m}$ by $% {\bf m}^{\prime \prime }\,$in the center generating function $S(x,{\bf m})$. Performing the mentioned replacements we will obtain: \begin{eqnarray} S(\xi ,{\bf m}^{\prime }) &=&S(\xi ,{\bf m})+\xi \wedge {\bf k-}\frac 12{\bf % m}\Gamma _1{\bf k-}\frac 14{\bf k}\Delta _1{\bf k} \label{scorper} \\ S(x,{\bf m}^{\prime \prime }) &=&S(x,{\bf m})-2x\wedge {\bf k-}\frac 12{\bf m% }\Gamma _2{\bf k-}\frac 14{\bf k}\Delta _2{\bf k,} \label{sxper} \end{eqnarray} where \begin{eqnarray} \Gamma _1 &=&\left[ \left( {\frak J}+\widetilde{{\frak J}}\right) {\cal M}+\left( \widetilde{{\frak J}}-{\frak J}\right) \right] \label{gama1} \\ \Delta _1 &=&\left[ \left( {\cal M}^t\widetilde{{\frak J}}{\cal M}+\widetilde{{\frak J}}% \right) +{\cal M}^t\left( \widetilde{{\frak J}}-{\frak J}\right) +\left( {\frak J}+% \widetilde{{\frak J}}\right) {\cal M}\right] \label{delta1} \\ \Gamma _2 &=&\left[ \left( {\frak J-}\widetilde{{\frak J}}\right) {\cal M}+\left( {\frak J}+\widetilde{{\frak J}}\right) \right] \label{gama2} \\ \Delta _2 &=&\left[ \left( {\cal M}^t\widetilde{{\frak J}}{\cal M}+\widetilde{{\frak J}}% \right) -{\cal M}^t\left( {\frak J}+\widetilde{{\frak J}}\right) -\left( {\frak J}-% \widetilde{{\frak J}}\right) {\cal M}\right] . \label{delta2} \end{eqnarray} In this way we can restrict ${\bf m}$ to integer component vectors that lie in one of the two fundamental parallelepipeds \begin{eqnarray} && \begin{array}{cc} \Diamond _\xi =({\cal M}-1)\Box & \qquad \mbox{for }S(\xi ,{\bf m}) \end{array} \label{para1} \\ && \begin{array}{cc} \Diamond _x=({\cal M}+1)\Box & \qquad \mbox{for }S(x,{\bf m}) \end{array} \label{para2} \end{eqnarray} where $\Box $ is the unit hypercube that denotes the $2L$-torus. Hence, the different orbits denoted by a given chord $\xi $ are given by all the integer ${\bf m}$ lying in $\Diamond _\xi $. The number of such orbits is independent of $\xi $, so taking $\xi =0$, we equate this to the number of fixed points $\tau _\xi $, i.e. \begin{equation} \tau ({\cal M})=|\det ({\cal M}-1)|=\frac{2^{2L}}{|\det ({\frak J}\beta -1)|}\equiv \tau _\xi . \label{d1} \end{equation} The different orbits that have the point $x$ as its center are denoted by all the integers ${\bf m}$ lying now in $\Diamond _x.$ The number of these orbits is given by the volume of $\Diamond _x$ which is \begin{equation} \tau (-{\cal M})=\left| \det ({\cal M}+1)\right| =\frac{2^{2L}}{|\det ({\frak J}B+1)|}% \equiv \tau _x. \label{gama} \end{equation} Note that the number of periodic points of period two is $\tau _\xi \tau _x.$ For the matrix ${\cal M}$ to represent a $2L$-cat map it must be symplectic, so that the map is area preserving, and ${\cal M}$ must have integer entries. Let us now translate both the conditions for the $B$ and $\beta $ matrices. The first condition implies that $B$ and $\beta $ are symmetric matrices, and any symmetric matrix is associated through (\ref{mbbeta}) to a symplectic matrix. The second condition restricts the $B$ and $\beta $ matrices to have rational entries. Indeed, following (\ref{jbm}) and(\ref{jbetam}) we will have \begin{eqnarray} B &=&\frac{\overline{B}}{\det ({\cal M}+1)}\equiv \pm \frac{\overline{B}}{\tau _x} \label{bbbar} \\ \beta &=&\frac{\overline{\beta }}{\det ({\cal M}-1)}=\pm \frac{\overline{\beta }}{% \tau _\xi } \label{betabar} \end{eqnarray} where $\overline{B}$ and $\overline{\beta }$ are symmetric matrices with integer entries and the denominators are defined by (\ref{d1}) and (\ref {gama}). It can happen that all the coefficients of the matrix $\overline{B}$ ( or $\overline{\beta }$ ) have a common factor that is not coprime with $% \tau _x$ (respectively $\tau _\xi $ ). Dividing by this common factor the fraction in (\ref{bbbar}) ( or in (\ref{betabar}) ) is reduced to the form \begin{eqnarray} B &=&\pm \frac{\overline{B}^{\prime }}{\tau _x^{\prime }} \label{bbbar1} \\ \beta &=&\pm \frac{\overline{\beta }^{\prime }}{\tau _\xi ^{\prime }}. \label{betabar1} \end{eqnarray} But not any symmetric $B$ or $\beta $ matrix with rational entries guarantees that the associated symplectic matrix will have integer elements. So we must find conditions on $B$ and $\beta $ for this to occur. The characterization of the matrices $B$ or $\beta $ requires that the corresponding transformation on the plane maps the points of an integer lattice among themselves. We will examine the case $L=1$, as the extension for many number of degrees of freedom follows easily. There are two fundamental chords corresponding to fixed points on the torus: \begin{equation} \xi _1=\left( \begin{array}{l} 1 \\ 0 \end{array} \right) \mbox{ \qquad and \qquad }\xi _2=\left( \begin{array}{l} 0 \\ 1 \end{array} \right) , \end{equation} leading to the fixed points: \begin{equation} x_j=\frac 12({\frak J}\beta + 1)\xi _{j\qquad ,}\mbox{ with }j=1,2. \end{equation} Of course, there is also $\xi _0=0$, but this ''plane fixed point'' makes no restriction on the torus map. For the transformation to be a cat map, all the corners of the fundamental parallelogram $\Box $ must be fixed points, so, for any integers $r$ and $s$, there are integers $m_1$ and $m_2$ such that: \begin{equation} m_1x_1+m_2x_2=\left( \begin{array}{l} r \\ s \end{array} \right) =\frac 12({\frak J}\beta + 1)\left( \begin{array}{l} {m_1} \\ {m_2} \end{array} \right) . \end{equation} This is true if only if $2({\frak J}\beta + 1)^{-1}$ has integer entries. This condition is general for any degrees of freedom. Similarly, we find that $2({\frak J}\beta - 1)^{-1}$ having integer entries is also a necessary and sufficient condition for the corresponding ${\cal M}^{-1} $ to define a cat map. But , if ${\cal M}$ defines a cat map, so does $-{\cal M}$ with the associated chord matrix $-B.$ Therefore, it is also a necessary and sufficient condition, for a center generating function to determine a cat map, that the associated center matrix $B$ have the property that $2({\frak J}B\pm 1)^{-1}$ be an integer matrix. Evidently we easily find a subclass of cat maps by restricting their Cayley parametrization to $B$ (or $ \beta$ ) matrices of the form (\ref{bbbar}) such that $ det(1\pm{\frak J}B)=\pm 1, \pm 2 $, then $2(1+{\frak J}B)^{-1}$ is an integer matrix. Although the conditions on the symmetric matrices $B$ or $\beta \,\,$to denote a cat map are not as trivial as the ones on the symplectic matrix ${\cal M},$ it is simpler to find rational symmetric matrices that fulfill the condition on $\beta $ and $B$ than to find integer symplectic matrices. The fact that $% B$ or $\beta $ are symmetric and of the form (\ref{bbbar}) allows us to find cat maps by sampling $\left[ (L)\times (2L+1)+1\right] $ integer numbers. Otherwise, to fulfill the condition (\ref{msimple}), needs a loop over $(2L)^2$ integer coefficients. To conclude this section, we verify the property of feline invariance for the chord and center generating functions. First, we note that, symplectic invariance in the plane \cite{ozrep} implies that under a symplectic coordinate transformation $x\rightarrow x^{\prime }={\cal N}x,$ $S(x,0)=S(x^{\prime },0)$ and $S(\xi ,0)=S(\xi ^{\prime },0),$ with $\xi ^{\prime }={\cal N}\xi .$ But it is also evident that the winding number ${\bf m}$ transforms in the same manner: ${\bf m}^{\prime }={\cal N}{\bf m.}$ As far as the $x$ dependent term in $% S(x,{\bf m}),$ we thus find that the effect of the feline transformation is merely that of substituting $B\rightarrow {\cal N}^tB{\cal N},$ and similarly the change in $S(\xi ,{\bf m})$ is obtained from $\beta \rightarrow {\cal N}^t\beta {\cal N}.$ The constant terms $f({\bf m})$ and $g({\bf m})$ in (\ref{sxf}) and (\ref{scorg}% ) are not invariant under a feline transformation in the form (\ref{gm}) that we have chosen to match reference \cite{keat1}, so that it is preferable to use (\ref{3.30}) and (\ref{3.30g}) when dealing with equivalence classes of cat maps. It is important to note that, unlike the symmetric matrices $B$ and $\beta $% , ${\frak J}B\rightarrow {\cal N}^{-1}{\frak J}B{\cal N}$ and ${\frak J}\beta \rightarrow {\cal N}^{-1}{\frak J}\beta {\cal N}$, under a similarity transformation ${\cal M}\rightarrow {\cal N}^{-1}{\cal M}{\cal N}.$ Therefor, the eigenvalues of ${\frak J}B$ and ${\frak J}\beta $ are feline invariant, just as those of ${\cal M}$, and can thus be used to classify cat maps. \section{Classification of classical cat maps} The periodic orbits of cat maps have been studied in great details by Percival and Vivaldi \cite{parcivivaldi} and also by Keating in \cite{keat1} for one degree of freedom and their results were recently extended to an arbitrary number of degrees of freedom \cite{greenman}. It is shown that a point on the unit $2L$-torus is periodic if and only if all its coordinates are rational and any grid of points with rational coordinates is invariant under the action of the map. From (\ref{mapa}) we can see that the periodic points $x_l$ of integer period $l$ are labeled by the winding numbers ${\bf % m,}$ so that \begin{equation} x_l=\left( \begin{array}{l} {p_l} \\ {q_l} \end{array} \right) =({\cal M}^l-1)^{-1}{\bf m=}({\frak J}\beta ^{\left( l\right) }-1)\frac{% {\bf m}}2. \label{xfix} \end{equation} Here $\beta ^{\left( l\right) }$ denotes the symmetric matrix associated to $% {\cal M}^l$ through (\ref{jbetam}). To have $x_l$ on the unit $2L$-hypercube $\Box $% , ${\bf m}$ must lie within the parallelepiped formed by the action of the matrix $({\cal M}^l-1)$ on $\Box .$ Hence, the number of integer points ${\bf m}$ is given by its hypervolume, so that the number of periodic points with period $l$ is \begin{equation} \tau ({\cal M}^l)=|\det ({\cal M}^l-1)|=\frac{2^{2L}}{|\det ({\frak J}\beta ^{\left( l\right) }-1)|}. \label{npfix} \end{equation} According to (\ref{xfix}) the periodic points of period $l$ form a lattice in phase space with rational coordinates. The motion of any point $x_{-}=x_1+\delta _{-}$ near a fixed point $x_1$ will be \begin{equation} {\cal M}x_{-}={\cal M}(x_1+\delta _{-})=x_1+{\cal M}\delta _{-}=x_1+\delta _{+}=x_{+}. \end{equation} To determine the character of such a motion we have to study the eigenvalues and eigenvectors of the matrix ${\cal M}$: $\,$ \begin{equation} \lambda _{\cal M}^k=\left| \lambda _{\cal M}^k\right| e^{i\theta _k} \end{equation} The modulus $\left| \lambda _{\cal M}^k\right| $ indicates that the motion is stretching while the argument $e^{i\theta _k}$ indicates rotation around the fixed point $x_1.$ For a symplectic matrix ${\cal M}$, if $\lambda _{\cal M}$ is an eigenvalue of ${\cal M}$ then $\lambda _{\cal M}^{*}$, $\frac 1{\lambda _{\cal M}}$ and $\frac 1{\lambda _{\cal M}^{*}}$ will also be eigenvalues of ${\cal M}.$ The classification of the eigenvalues is possible in either the ${\cal M},B,$ or $% \beta $ descriptions. Using (\ref{mbbeta}) we obtain the relation with the eigenvalues of ${\frak J}B$ and ${\frak J}\beta $ denoted respectively as $% \lambda _{{\frak J}B}$ and $\lambda _{{\frak J}\beta },$% \begin{equation} \lambda _{\cal M}=\frac{\left( 1-\lambda _{{\frak J}B}\right) }{\left( 1+\lambda _{% {\frak J}B}\right) }=\frac{\left( \lambda _{{\frak J}\beta }+1\right) }{% \left( \lambda _{{\frak J}\beta }-1\right) } \label{lmbbeta} \end{equation} and inversely: \begin{eqnarray} \lambda _{{\frak J}B} &=&\frac{\left( 1-\lambda _{\cal M}\right) }{\left( 1+\lambda _{\cal M}\right) } \label{ljbm} \\ \lambda _{{\frak J}\beta } &=&\frac{\left( \lambda _{\cal M}+1\right) }{\left( \lambda _{\cal M}-1\right) }=-\frac 1{\lambda _{{\frak J}B}}. \label{ljbetam} \end{eqnarray} In this way, if $\lambda _{{\frak J}B}$ is an eigenvalue of ${\frak J}B,$ then $\lambda _{{\frak J}B}^{*}$, $-\lambda _{{\frak J}B}$ and $-\lambda _{% {\frak J}B}^{*}$ will also be. In the same way, if $\lambda _{{\frak J}\beta }$ is an eigenvalue of ${\frak J}\beta ,$ then $\lambda _{{\frak J}\beta }^{*}$, $-\lambda _{{\frak J}\beta }$ and $-\lambda _{{\frak J}\beta }^{*}$ also are. For cat maps with $L=2$ the matrices ${\cal M},B,$ and $\beta $ will be $4\times 4$ and we then have the following generic cases for the eigenvalues \begin{enumerate} \item Elliptic: there is a conjugate pair of $\lambda _{\cal M}$ both on the unit circle and conjugate pairs of purely imaginary $\lambda _{{\frak J}B}$ and $% \lambda _{{\frak J}\beta }.$ \item Hyperbolic: there is a pair $(\lambda _{\cal M},\frac 1{\lambda _{\cal M}})$ on the real axis and pairs $(\lambda _{{\frak J}B},-\lambda _{{\frak J}B})$ and $% (\lambda _{{\frak J}\beta },-\lambda _{{\frak J}\beta })$ on the real axis. \item Parabolic: there are degenerate eigenvalues $\lambda _{\cal M}=\pm 1$. Then for $\lambda _{\cal M}=1,$ $\beta $ is singular and $\lambda _{{\frak J}B}=0$; for $% \lambda _{\cal M}=-1,$ $B$ is singular and $\lambda _{{\frak J}\beta }=0$. \item Mixed: each pair belongs to a different one of the above categories. \item Loxodromic: the eigenvalues of ${\cal M},B,$ or $\beta $ matrices form quartets of complex eigenvectors of the form $(\lambda _{\cal M},\frac 1{\lambda _{\cal M}},\lambda _{\cal M}^{*},\frac 1{\lambda _{\cal M}^{*}})\,$ for the ${\cal M}$ matrix and $% (\lambda _{{\frak J}S},-\lambda _{{\frak J}S},\lambda _{{\frak J}% S}^{*},-\lambda _{{\frak J}S}^{*})\,$ for $S$ being one of the symmetric matrices $B$ or $\beta .$ \end{enumerate} The first three cases arise also for one degree of freedom cat maps. The case $L=2$ is the lowest number of degrees of freedom where not only elliptic, parabolic and hyperbolic fixed points will appear, but also loxodromic ones. For more degrees of freedom no new cases will occur; there is only a greater variety of mixed cases. Nongeneric possibilities arise for any dimension for continuous families of systems \cite{quinze}, but we do not know if there exists any corresponding cat map. In the general case, the motion is the composition of stretching and rotation, but if the angles of the rotation are of the form $\theta _k=\frac ij2\pi ,$ after $j\,$ applications of the map, the matrix ${\cal M}^j$ has only real eigenvalues that come in pairs $\left| \lambda _{\cal M}^k\right| ^j$ and $% \left| \lambda _{\cal M}^k\right| ^{-j}$. The dynamics has then an ignorable coordinate; the angles are constants of the motion for ${\cal M}^j.$ For one degree of freedom systems, where there are only two eigenvalues, $\lambda _{\cal M}$ and $% \lambda _{\cal M}^{*},$\ the condition \begin{equation} Tr\left( {\cal M}\right) =\lambda _{\cal M}+\lambda _{\cal M}^{*}=\mbox{integer} \end{equation} implies that in the elliptic case only angles of the form $\theta _k=\frac ij\pi $ with $j=2$ or $3$ are allowed. This is an example of a rational restriction on $\theta _k$, that is, ${\cal M}^j$ will be the identity map. This result is in accordance with Ma\~{n}e's theorem \cite{mane} that two-dimensional symplectomorphisms are either Anosov or they have zero entropy. Nonetheless, irrational rotation angles do exist in the loxodromic or mixed cases for $L>1$. We now turn our attention to the classification of four-dimensional cat maps. Recalling that the characteristic polynomial for any $k\times k$ matrix $A$ is \begin{equation} P_A(\lambda )=\det (A-\lambda )=\sum_{n=1}^k\alpha _n\lambda ^{k-n}, \end{equation} where the $\alpha _n$ coefficients are given by the recurrence relation, \begin{eqnarray} \alpha _0 &=&1, \\ \alpha _n &=&-\frac 1n\sum_{i=1}^n\alpha _{n-i}a_i\qquad \mbox{ with }% a_i=Tr(A^i), \end{eqnarray} we obtain that for any symmetric matrix $B$ (i.e. $B$ and $\beta $ ), \begin{equation} P_{{\frak J}B}(\lambda )=P_{{\frak J}B}(-\lambda ) \label{psym} \end{equation} so that \begin{equation} P_{{\frak J}B}(\lambda )=\lambda ^4-\frac 12b_2\lambda ^2+\det B \label{pbl} \end{equation} with $b_2=Tr\left[ \left( {\frak J}B\right) ^2\right] ,($ note that $\det {\frak J}B=\det B)$. There is a similar expression for $P_{{\frak J}\beta }(\lambda )$, required when the center representation is singular, \begin{equation} P_{{\frak J}\beta }(\lambda )=\lambda ^4-\frac 12\beta _2\lambda ^2+\det \beta . \label{pbetal} \end{equation} where $\beta _2=Tr\left[ \left( {\frak J}\beta \right) ^2\right] .$ For the symplectic case, we obtain \begin{equation} P_{\cal M}(\lambda )=\lambda ^4-Tr({\cal M})\lambda +\frac 12\left[ Tr({\cal M}^2)-Tr({\cal M})^2\right] \lambda ^2-Tr({\cal M})\lambda ^3+1 \label{pml} \end{equation} which is harder to analyze, so we will perform the classification of the different behaviors using ${\frak J}B$ or ${\frak J}\beta $. Solving $P_{% {\frak J}B}(\lambda )=0$ using (\ref{pbl}), leads to \begin{equation} \lambda _{{\frak J}B}=\pm \sqrt{\frac{b_2}4\pm \sqrt{\left( \frac{b_2}% 4\right) ^2-\det B}}, \end{equation} whereas \begin{equation} \lambda _{{\frak J}\beta }=\pm \sqrt{\frac{\beta _2}4\pm \sqrt{\left( \frac{% \beta _2}4\right) ^2-\det \beta }}. \end{equation} We can now classify the different behaviors according to the feline invariants of the matrix ${\frak J}B\,$ or ${\frak J}\beta $, the eigenvalues of the symplectic matrix ${\cal M}$ being obtained with the help of (% \ref{lmbbeta}). The loxodromic behavior corresponding to four complex eigenvalues will appear if the square root term $\left[ \left( \frac{b_2}% 4\right) ^2-\det B\right] $ inside the square root is negative. That is, the loxodromic behavior will appear only if \begin{equation} \det B>\left( \frac{b_2}4\right) ^2\qquad \mbox{ or equivalently }\qquad \det \beta >\left( \frac{\beta _2}4\right) ^2. \label{loxcond} \end{equation} In fig~\ref{fig.1} we find a complete classification of the different types of behavior according to the feline invariants of ${\frak J}\beta $ and the same arises for the invariants of ${\frak J}B.$ These invariants also allow us to obtain $\tau _\xi ,\,$the number of orbits for any chord $% \xi ,$% \begin{equation} \tau _\xi =\left| \frac{2^4}{P_{{\frak J}\beta }(1)}\right| =\left| \frac{% 2^4\det B}{P_{{\frak J}B}(1)}\right| , \end{equation} or $\tau _x$ the number of orbits centered on $x$ as \begin{equation} \tau _x=\left| \frac{2^4}{P_{{\frak J}B}(1)}\right| =\left| \frac{2^4\det \beta }{P_{{\frak J}\beta }(1)}\right| . \end{equation} \begin{figure}[tbp] \centerline {\epsfxsize=5in \epsffile{fig1.eps} } \caption{ A classification of the different cat maps for two degrees of freedom according to the feline invariants of the ${\frak J}\beta $ matrix. The parabolic boundary is $\det \beta =\left( \frac{\beta _2}4\right) ^2.$ A similar picture exist for the ${\frak J}B$ matrix. } \label{fig.1} \end{figure} \subsection*{Examples} We here show some examples of different types of cat maps. The examples are such that the $\beta $ matrix, that denotes the chord representation, has integer elements. \subsubsection*{ The Hannay and Berry Cat} It is important to see that the first cat map to be quantized ( the Hannay and Berry cat map) \cite{hanay} can also be treated with the formalism described here . For a $2\times 2$ matrix, the characteristic polynomial reads, \begin{equation} P_{{\frak J}\beta }(\lambda )=\lambda ^2-\frac 12\beta _2. \end{equation} For the symplectic matrix \begin{equation} {\cal M}_{hb}=\left[ \begin{array}{cc} 2 & 1 \\ 3 & 2 \end{array} \right] \mbox{, with the associated symmetric matrix }\beta _{hb}=\left[ \begin{array}{cc} 3 & 0 \\ 0 & -1 \end{array} \right] , \label{mhb} \end{equation} so that \[ \beta _2=Tr\left[ \left( {\frak J}\beta _{hb}\right) ^2\right] =6, \] the number of fixed points is \[ \tau _\xi ^{hb}=\left| \frac{2^2}{P_{{\frak J}\beta }(\lambda )}\right| =2 \] and the eigenvalues of the matrix ${\cal M}_{hb}$ are \begin{eqnarray*} \lambda _1^{hb} &=&2+\sqrt{3} \\ \lambda _2^{hb} &=&2-\sqrt{3}. \end{eqnarray*} The map is then hyperbolic. \subsubsection*{ The double hyperbolic case} Let us now study cat maps that have two pairs of eigenvalues both in the real axis, that is, maps belonging to the first quadrant in figure \ref{fig.1}% . The $\beta _{hh}\,$ matrix below describes this case, the associated symplectic matrix being ${\cal M}_{hh}$ : \begin{equation} \beta _{hh}=\left[ \begin{array}{cccc} 0 & 2 & 1 & 2 \\ 2 & 0 & 2 & 1 \\ 1 & 2 & 0 & 0 \\ 2 & 1 & 0 & 1 \end{array} \right] \quad \mbox{and }\quad {\cal M}_{hh}=\left[ \begin{array}{cccc} 2 & -2 & -1 & 0 \\ -2 & 3 & 1 & 0 \\ -1 & 2 & 2 & 1 \\ 2 & -2 & 0 & 1 \end{array} \right] . \label{mhh} \end{equation} The invariants are \[ \det \beta _{hh}=17\quad \mbox{and }\quad \beta _2=Tr\left[ \left( {\frak J}% \beta _{hh}\right) ^2\right] =20. \] Thus, the number of fixed points $\tau _\xi ^{hh}=2$ and the eigenvalues of the symplectic matrix ${\cal M}$ are \begin{eqnarray*} \lambda _1^{hh} &=&2.112388 \\ \lambda _2^{hh} &=&0.4739 \\ \lambda _3^{hh} &=&5.22 \\ \lambda _4^{hh} &=&0.1914. \end{eqnarray*} \subsubsection*{ The mixed case} We consider here the case where the eigenvalues of the map are of mixed nature, i.e. there is a couple of real eigenvalues and a complex conjugated pair on the unit circle. This corresponds to $\beta $ matrices that belong to the third and fourth quadrant in figure ~\ref{fig.1}. One matrix of this kind is $\beta _{eh1}$, the associated symplectic matrix being ${\cal M}_{eh1}:$ \begin{equation} \beta _{eh1}=\left[ \begin{array}{cccc} 0 & 2 & 1 & 2 \\ 2 & 0 & 2 & 1 \\ 1 & 2 & 0 & 2 \\ 2 & 1 & 2 & 0 \end{array} \right] \quad \mbox{and }\quad {\cal M}_{eh1}=\left[ \begin{array}{cccc} 0 & 0 & -1 & 0 \\ 0 & 0 & 0 & -1 \\ 1 & 0 & 2 & 1 \\ 0 & 1 & 1 & 2 \end{array} \right] . \label{meh1} \end{equation} The invariants are \[ \det \beta _{eh1}=-15\quad \mbox{and }\quad \beta _2=Tr\left[ \left( {\frak J% }\beta _{eh1}\right) ^2\right] =4. \] Thus, the number of fixed points $\tau _\xi ^{eh1}=1$ and the eigenvalues of the symplectic matrix ${\cal M}$ are \begin{eqnarray*} \lambda _1^{eh1} &=&2.6180 \\ \lambda _2^{eh1} &=&0.381966 \\ \lambda _3^{eh1} &=&\exp \left( i\frac{2\pi }6\right) \\ \lambda _4^{eh1} &=&\exp \left( -i\frac{2\pi }6\right) . \end{eqnarray*} This example shows rotation angles that are fractional multiples of $\pi $, for which the dynamics are equivalent to that of hyperbolic systems with one degree of freedom . Another example of a mixed system is given by the matrix $\beta _{eh2}$, the associated symplectic matrix being ${\cal M}_{eh2}:$ \begin{equation} \beta _{eh2}=\left[ \begin{array}{cccc} 0 & 2 & 1 & 2 \\ 2 & 0 & 2 & 1 \\ 1 & 2 & 0 & 2 \\ 2 & 1 & 2 & 1 \end{array} \right] \quad \mbox{and }\quad {\cal M}_{eh2}=\left[ \begin{array}{cccc} 0 & 0 & -1 & 0 \\ 0 & -1 & -1 & -2 \\ 1 & 0 & 2 & 1 \\ 0 & 2 & 2 & 3 \end{array} \right] . \label{meh2} \end{equation} Now we have \begin{equation} \det \beta _{eh2}=-7\quad \mbox{and }\quad \beta _2=Tr\left[ \left( {\frak J}% \beta _{eh2}\right) ^2\right] =4. \end{equation} Thus, the number of fixed points $\tau _\xi ^{eh2}=2$ and the eigenvalues of the symplectic matrix ${\cal M}$ are \begin{eqnarray*} \lambda _1^{eh2} &=&3.0906578 \\ \lambda _2^{eh2} &=&0.32355571 \\ \lambda _3^{eh2} &=&\exp \left( i1.27354496\right) \\ \lambda _4^{eh2} &=&\exp \left( -i1.27354496\right) . \end{eqnarray*} In this example there are no rotations with angles that are a fraction of $% \pi ,$ the dynamics will then be ergodic and mixing in the whole phase space. \subsubsection*{ The loxodromic case} We choose now a $\beta $ matrix that belongs to the loxodromic region in Fig ~\ref{fig.1}, for example $\beta _{lox}$, whose associated symplectic matrix being ${\cal M}_{eh1}:$ \begin{equation} \beta _{lox}=\left[ \begin{array}{cccc} 0 & 2 & 1 & 0 \\ 2 & 0 & 2 & 1 \\ 1 & 2 & 1 & 1 \\ 0 & 1 & 1 & 1 \end{array} \right] \quad \mbox{and }\quad {\cal M}_{lox}=\left[ \begin{array}{cccc} 0 & 1 & 0 & 0 \\ 0 & 1 & 1 & 0 \\ 1 & -1 & 1 & 1 \\ -1 & -1 & -2 & 0 \end{array} \right] \label{mlox} \end{equation} whose invariants are \begin{equation} \det \beta _{lox}=5\quad \mbox{and }\quad \beta _2=Tr\left[ \left( {\frak J}% \beta _{lox}\right) ^2\right] =-4. \nonumber \end{equation} Thus, the number of fixed points $\tau _\xi ^{lox}=2$ and the eigenvalues of the symplectic matrix ${\cal M}$ are \begin{eqnarray*} \lambda _1^{lox} &=&1.7000157\exp \left( i1.1185178\right) \\ \lambda _2^{lox} &=&1.7000157\exp \left( -i1.1185178\right) \\ \lambda _3^{lox} &=&0.5882298\exp \left( i1.1185178\right) \\ \lambda _4^{lox} &=&0.5882298\exp \left( -i1.1185178\right) . \end{eqnarray*} There are no rational rotations, so the dynamics is ergodic and mixing in the whole phase space. \subsubsection*{The double elliptic case} In this case, the eigenvalues are characterized by two rotation angles. An interesting feature is that the condition that the ${\cal M}$ matrix have integer elements, prevents either angle from being an irrational multiple of $\pi .$ This restricts the dynamics to the trivial integrability observed for one degree of freedom systems. In conclusion, any symplectomorphism on the $4$-torus belongs to one of the following cases: \begin{enumerate} \item Ergodic on the full phase space (a manifold of 2 degrees of freedom). For the double hyperbolic, mixed or loxodromic cases with irrational rotation angles. \item Ergodic in a $2$-dimensional manifold . We have a constant of the motion, then the ergodicity is restricted to a low dimensional manifold (a manifold with $L=1).$ For the mixed or loxodromic cases with rational rotation angles. \item A root of unity: that is, we have two constants of the motion. All orbits are periodic and have the same period for this double elliptic cases. \end{enumerate} \section{Simple Quantum Cat maps} \setcounter{equation}{0} Quantum dynamics is characterized by a unitary evolution operator, or propagator. It is possible to obtain the center and chord representation of an operator on the torus from its counterpart on the plane. This will be the way that we quantize cat maps, leaving the general construction for the next section. In some cases the quantum propagator denoting cat maps in the center or chord representation acquires simple expressions; in this section we will treat these special cases. More details about torus Hilbert space and its center and chord representations based on reference \cite{opetor} are available in Appendix A. It is well known that torus quantization implies that the Hilbert space $% \left[ {\cal H}_N^\chi \right] ^L$ associated to the $2L$-torus has finite dimension $N^L$ and is characterized by a vector Floquet parameter $\chi =(\chi _p,\chi _q)$ whose components are real numbers belonging to $\left[ 0,1\right] .$ The fact that $\left[ {\cal H}_N^\chi \right] ^L$ has finite dimension, implies that position and momentum eigenstates can only take on a set of discrete values that form a discrete lattice called {\it quantum phase space} (QPS). Any point $x$ in this QPS has coordinates, \begin{equation} x=\left( \begin{array}{c} p_m \\ q_n \end{array} \right) =\frac 1N\left( \begin{array}{c} m+\chi _p \\ n+\chi _q \end{array} \right) . \end{equation} The center and chord representations are based on translation and reflection operators on this QPS as we show in \cite{opetor}. Chords are of the form \begin{equation} \xi _{r,s}=\frac 1N\left( \begin{array}{c} r \\ s \end{array} \right) =\frac 1N\bar{\xi}, \end{equation} with $r$ and $s$ integer numbers and $\bar{\xi}=\left( \begin{array}{c} r \\ s \end{array} \right) $, while the center points $x_{a,b}$ are labeled by half-integer numbers $a$ and $b,$ \begin{equation} x_{a,b}=\frac 1N\left( \begin{array}{c} a+\chi _p \\ b+\chi _q \end{array} \right) . \end{equation} At a first stage we restrict our attention to those maps that can be quantized on the Floquet parameters $\chi =(0,0).$ As we will see in section 5 this implies that the matrix ${\cal M}$ must satisfy \begin{equation} \sum_{j=1}^L{\cal M}_{i,j}{\cal M}_{i,j+L}=\mbox{even integer for all }i. \label{Mquanti1} \end{equation} As we will show in section 5, if $2\tau _\xi ^{\prime }$, defined in (\ref {betabar1}), and $N$ are coprime numbers, a complete representation of the propagator is obtained having the symbol on a lattice of chords $\Xi $ such that \begin{equation} \Xi =\xi +{\bf n=}\frac{2\tau _\xi ^{\prime }}N\overline{\Xi }, \label{corN1} \end{equation} where the components of $\overline{\Xi }$ are integer numbers up to $N.$ So for any chord $\xi $ there is an equivalent chord $\Xi .$ For the allowed values of $N$, the propagator for cat maps in the chord representation takes the form \begin{equation} {\bf U}_{\cal M}(\Xi )=2^L\left( \tau _\xi \right) ^{-\frac 32}e^{-i2\pi N\left[ \frac 14\Xi \beta \Xi \right] }\sum_{{\bf m\in \Diamond }_\xi }e^{-i2\pi N\frac 14{\bf m}(\beta -\widetilde{{\frak J}}){\bf m}}, \label{ugcorimp} \end{equation} where $\tau _\xi $ is the number of fixed points of the classical map. For $% \beta $ matrices that fulfill the feline conditions, the symbol ${\bf U}% _{\cal M}^0(\Xi )$ must represent an unitary operator. In that case (\ref{unitcor}) shows that it must has the form \begin{equation} {\bf U}_{\cal M}(\Xi )=\frac{e^{i\varphi _N({\cal M})}}{\sqrt{N}^L}e^{-i2\pi N\left[ \frac 14\Xi \beta \Xi \right] }, \end{equation} which restricts \begin{equation} \frac{e^{i\varphi _N({\cal M})}}{\sqrt{N}^L}=2^L\left( \tau _\xi \right) ^{-\frac 32}\sum_{{\bf m\in \Diamond }_\xi }e^{-i2\pi N\frac 14{\bf m}(\beta -% \widetilde{{\frak J}}){\bf m}}. \end{equation} At first sight, the phase $\varphi _N({\cal M})$ is only an unimportant global phase factor, but the interference of the different $\varphi _N({\cal M}^l)$ for the different powers $l$ of the map will have a crucial importance for the density of states. As $\xi $ and $\Xi $ are equivalent chords, the symbol ${\bf U}_{\cal M}(\xi )\,$ and ${\bf U}_{\cal M}(\Xi )$ are related by symmetry relations (\ref{eq:Acorsim}) so that \begin{equation} {\bf U}_{\cal M}(\xi )=\frac{e^{i\varphi _N({\cal M})}}{\sqrt{N}^L}e^{-i2\pi N\left[ S(\xi ,{\bf n})\right] }, \label{ucors} \end{equation} where $S(\xi ,{\bf n})$ is the action of the classical orbit whose chord is $% \xi $ and that performs ${\bf n}$ loops around the torus as defined in (\ref {scor}). If we chose the symplectically invariant form (\ref{3.30g}) for the $\xi $% -independent part of the chord generating function, instead of (\ref{gm}), we would have in (\ref{ucors}) a supplementary phase factor $e^{i2\pi N\frac 14{\bf n}\widetilde{{\frak J}}{\bf n}}=e^{i\gamma _{{\bf n}}}$ with $\gamma _{{\bf n}}$ a "Maslov index" for the orbit. This observation is true for all the following quantum theory. In the case of the center representation, for $2\tau _x^{\prime }\,,$ defined in (\ref{bbbar1}), and $N$ coprime numbers, a complete representation of the propagator is obtained by performing a transformation to center points $X$ that are integer multiples of $\frac{\tau _x^{\prime }}N$, \begin{equation} X=x+\frac 12{\bf j=}\frac{\tau _x^{\prime }}N\overline{X}, \label{xN1} \end{equation} where the components of $\overline{X}$ are integer numbers up to $N$ \cite {opetor}$.$ On these points the center representation of the propagator takes the form \begin{eqnarray} {\bf U}_{\cal M}(X) &=&2^L\tau _x^{-\frac 32}e^{i2\pi N\left[ XBX\right] }\sum_{% {\bf m\in \Diamond }_x}e^{i2\pi N\frac 14{\bf m}(B+\widetilde{{\frak J}})% {\bf m}} \label{ugxd} \\ &=&e^{i\varphi _N^{\prime }({\cal M})}e^{i2\pi N\left[ XBX\right] }, \end{eqnarray} where the last equality is obtained by imposing the unitarity of ${\bf \hat{U% }}_{\cal M}$ and using (\ref{unitx}). Hence, we define the angle $\varphi _N^{\prime }({\cal M})$ so \begin{equation} e^{i\varphi _N^{\prime }({\cal M})}=2^L\tau _x^{-\frac 32}\sum_{{\bf m\in \Diamond }% _x}e^{i2\pi N\frac 14{\bf m}(B+\widetilde{{\frak J}}){\bf m}}. \end{equation} From the symmetry relations (\ref{eq:Acensim}), we find that the symbols on the original points $x$ are \begin{equation} {\bf U}_{\cal M}(x)=e^{i\varphi _N^{\prime }({\cal M})}e^{i2\pi N\left[ S(x,{\bf j}% )\right] }, \end{equation} where here $S(x,{\bf j})\,$ is the center generating function, defined in (% \ref{sx}), on a center point $x$ for an orbit performing ${\bf j}$ loops. The cases above are then special cases where the propagator on the torus has the same form as its equivalent on the plane; thus, they are ideally suited for the comparison of classical and quantum motion. To obtain the more familiar position representation of the propagator from its chord representation, we use (\ref{eq:AQcor}), \begin{equation} {\bf U}_{\cal M}{\bf (q}_m,{\bf q}_n)=\frac{e^{i\varphi _N({\cal M})}}{(N)^{\frac{3L}2}}% \sum_{\xi _p=0}^{N-1}\exp \left\{ -i2\pi N\left[ S(\xi _{p,m-n},{\bf n)+}% \frac{q_m+q_n}2\xi _p\right] \right\} , \end{equation} While (\ref{eq:AQx}) allows us to obtain the position representation from the center one: \begin{equation} {\bf U}_{\cal M}{\bf (q}_m,{\bf q}_n)=\frac{e^{i\varphi _N^{\prime }({\cal M})}}{N^L}% \sum_{x_p=0}^{\frac{N-1}2}\exp \left\{ i2\pi N\left[ S(x_{p,\frac{m+n}2},% {\bf j)+}\left( q_m-q_n\right) x_p\right] \right\} . \label{eq:UQx} \end{equation} For the case of one degree of freedom, this leads to the Hannay and Berry propagator. As we will see in section 5, condition (\ref{Mquanti1})\ is preserved for the different powers of the map, hence, if a map is quantizable on the Floquet parameters $\chi =(0,0),$ all its powers also are. Furthermore, the propagator for ${\cal M}^l$ has the form \begin{equation} {\bf U}_{\cal M}(\Xi )=\frac{e^{i\varphi _N({\cal M})}}{\sqrt{N}^L}e^{-i2\pi N\left[ \frac 14\Xi \beta ^{(l)}\Xi \right] }, \label{ugcorl} \end{equation} in the chord representation, where $\beta ^{(l)}$ denotes the symmetric matrix associated to ${\cal M}^l$ through (\ref{jbetam}) and $\Xi $ are points on a lattice of the form \begin{equation} \Xi {\bf =}\frac{\tau ({\cal M}^l)}N\overline{\Xi }. \end{equation} As we discussed in section 2, lattices of rational points are invariant under classical cat maps. The denominator $g$ of these rational points can then be used to label the lattice, i.e. there exists a minimal period $l_g$ under which all the points on the lattice are fixed for ${\cal M}^{l_g}.\,$In other words, ${\cal M}^{l_g},$ restricted to the mentioned lattice, is the identity. We will call $l_g$ the classical periodicity function of the lattice. As we can see in the appendix, torus quantization is performed on a lattice of rational points whose denominator is $N$. Under these considerations we find that the quantum propagator is also periodic. It is now possible to define conditions on the matrix ${\cal M}$ and the dimension of the Hilbert space $N\,$ under which $\widehat{{\bf U}}_{\cal M}\,$ reduces to the identity operator, i.e. \begin{equation} \widehat{{\bf U}}_{\cal M}=\widehat{{\bf 1}}_Ne^{i\phi }. \end{equation} This is best seen in the center representation, where we have (\ref{unox}) \begin{equation} {\bf U}_{\cal M}(x)={\bf 1}_N(x)e^{i\phi }=e^{i\phi }f_N(x), \label{uxident} \end{equation} with $f_N(x)$ defined in (\ref{eq:trRT}). For the case where $N$ is an odd integer, we can perform the quantization on points $X$ that are integer multiples of $\frac{\tau _x^{\prime }}N$ , as we already discussed$.$ We can see that if the $B$ matrix has all its elements that are multiples of $\frac N{\tau _x^{\prime }},$ then the propagator given by (\ref{ugxd})\ has the form (\ref{uxident}), where $f_N(X)=1$ and $\phi =\varphi _N^{\prime }({\cal M})$. Hence, the $B$ matrix denotes the identity if all its coefficients are multiples of $\frac N{\tau _x^{\prime }}.$ This implies through (\ref{mbbeta}% )\ that the ${\cal M}$ matrix must have the form \begin{equation} {\cal M}={\bf 1\quad }\mbox{mod}(N), \end{equation} in agreement with reference \cite{hanay} for the $L=1$ case. For any matrix ${\cal M}\,$ and for a given odd value of $N$, there is some smallest integer $k(N)$ such that \begin{equation} {\cal M}^{k(N)}={\bf 1\quad }\mbox{mod}(N) \label{qpf} \end{equation} and in this way \begin{equation} \left[ \widehat{{\bf U}}_{\cal M}^0\right] ^{k(N)}={\bf 1}e^{i\phi (N)}. \end{equation} In this sense, we may call the quantum propagator periodic with a period equal to $k(N)\,$ that we then call the {\it quantum period function} (QPF) although it is completely determined by the classical map through (\ref{qpf}% ). Note that $k(N)=l_N$, i.e., the quantum and classical period function coincide for $N$ an odd number, as we would expect. It is now easy to see that the $N^L\,$ eigenangles $\theta _m$ of the unitary propagator $\widehat{% {\bf U}}_{\cal M}$ are then restricted to lie on the $k(N)$ possible sites \begin{equation} \left\{ \alpha _j=\left[ \frac{2j\pi +\phi (N)}{k(N)}\right] \right\} ,1\leq j\leq k(N). \label{alfa} \end{equation} The eigenangle spectrum may be related to the traces of the propagator \begin{eqnarray} {\bf Tr}\left[ \left( \widehat{{\bf U}}_{\cal M}\right) ^l\right] &=&\sum_{m=1}^{N^L}e^{il\theta _m} \\ &=&\sum_{j=1}^{k(N)}d_je^{il\alpha _j} \end{eqnarray} where $d_j$ is the degeneracy of the $j$th site defined by (\ref{alfa}). The density of states, \begin{equation} \rho (\theta )=\sum_{i=l}^{N^L}\sum_{l=-\infty }^\infty \delta (\theta -\theta _i+2\pi l) \end{equation} is clearly invariant under $\theta \rightarrow \theta +2\pi .$ Use the Poisson summation formula, leads to the trace formula, \begin{equation} \rho (\theta )=\frac{N^L}{2\pi }+\frac 1\pi \mbox{Re}\sum_{l=1}^\infty {\bf % Tr}\left[ \left( \widehat{{\bf U}}_{\cal M}\right) ^l\right] e^{-il\theta }, \label{romapa} \end{equation} which holds for all maps on the $2L$-torus. Now, for the cat map, the fact that the propagator has periodicity $k(N)$ means that the density of states can also be written in the form \cite{keat2} \begin{equation} \rho (\theta )=\sum_{l=1}^{k(N)}{\bf Tr}\left[ \left( \widehat{{\bf U}}% _{\cal M}\right) ^l\right] e^{-il\theta }\sum_{j=-\infty }^\infty \delta (\phi (N)+2\pi j-k(N)\theta ). \label{rogato} \end{equation} That is, the eigenangles are restricted to the sites $\alpha _j$ in (\ref {alfa}), with the degeneracy at the $j$th site being given by \begin{equation} d_j=\frac 1{k(N)}\sum_{l=1}^{k(N)}{\bf Tr}\left[ \left( \widehat{{\bf U}}% _{\cal M}\right) ^l\right] e^{-il\alpha _j}. \end{equation} This last equation then leads to simple expressions for several important properties of the eigenangle distribution; for example \begin{equation} \sum_{j=1}^{k(N)}d_j^2=\frac 1{k(N)}\sum_{l=1}^{k(N)}\left| {\bf Tr}\left[ \left( \widehat{{\bf U}}_{\cal M}\right) ^l\right] \right| ^2. \end{equation} It is important to note that the $\delta $-functions appear explicitly in (% \ref{rogato}), because the propagator is periodic. The different powers of the quantum map in (\ref{rogato}) contribute only to the finite degeneracies, not to the $\delta $-function form of the trace formula. Moreover, for all finite $N$, only a finite number of these powers are required to determine the degeneracy at every available site, and hence give the whole spectrum. However, $k(N)\rightarrow \infty $ as $N\rightarrow \infty $, thus more terms are required to determine the spectrum as the semiclassical limit is approached. We now take the trace of (\ref{ugcorl}), using (\ref{TRA}), to obtain \begin{eqnarray} {\bf Tr}\left[ \left( \widehat{{\bf U}}_{\cal M}\right) ^l\right] &=&{\bf U}% _{{\cal M}^l}(\xi =0) \\ &=&2^L\left( \tau ({\cal M}^l)\right) ^{-\frac 32}\sum_{{\bf m\in \Diamond }_\xi ^l}\exp \left[ -i2\pi N\frac 14{\bf m}(\beta ^{(l)}-\widetilde{{\frak J}})% {\bf m}\right] =\frac{e^{i\varphi _N({\cal M}^l)}}{\sqrt{N}^L}, \end{eqnarray} where we recognize the action of the periodic orbits in the summation exponent, so \begin{eqnarray} {\bf Tr}\left[ \left( \widehat{{\bf U}}_{\cal M}\right) ^l\right] &=&2^L\left| \det ({\cal M}^l-1)\right| ^{-\frac 32}\sum_{{\bf m\in \Diamond }_\xi ^l}e^{-i2\pi N\left[ S_f^l({\bf m})\right] } \label{guttrf} \\ &=&\frac{e^{i\varphi _N({\cal M}^l)}}{\sqrt{N}^L}. \label{inter} \end{eqnarray} Note that ${\bf m\in \Diamond }_\xi ^l$ implies that we are summing over the periodic orbits of period $l$. Equation (\ref{guttrf}) represents the {\it % Gutzwiller-Tabor trace formula} \cite{tabor} for the cat map, though we must note that, instead of being a semiclassical approximation, it is exact in this case. We must also note that the expression (\ref{inter}) for the traces of the propagator implies that in the expression (\ref{rogato}) for the density of states, we have interference of the different phase factors $% e^{i\varphi _N({\cal M}^l)}$ for the different powers $l$ of the map. \subsection*{Examples :} We will now study some quantum features of the classical examples studied in section 3. All of them fulfill condition (\ref{Mquanti1}), so that their symmetric $\beta $ matrix has integer entries. Then quantization can be performed for all odd values of $N,$ for which the quantum propagator will have the form \begin{equation} {\bf U}_{\cal M}(\Xi )=\sqrt{\frac 1{N^L}}e^{-i2\pi N\left[ \frac 14\Xi \beta \Xi \right] } \end{equation} in the chord representation. In this case, the chords $\Xi $, defined in (% \ref{corN}), form a lattice of spacing $\frac 2N.$ In the following, we will study the quantum period function for these maps. We take cat maps of two degrees of freedom already studied in section 2. These include all the possible classical behaviors and we will study their effect on the QPF defined in (\ref{qpf}). \subsubsection*{ The Hannay and Berry cat map} To compare the different types of behavior that occur for two degrees of freedom, it is important to present the already known QPF for the Hannay and Berry cat map whose symplectic matrix is ${\cal M}_{hb}$ defined in (\ref{mhb}). The QPF\ is shown in figure ~\ref{fig.2}, where we can then see that, although it has an irregular behavior as a whole, most of the points lie in families of straight lines which admit a maximum slope. Indeed, this indicates that there exists an increasing sequence of primes $p$ such that $% \frac p{k(p)}<C$ for some $C$ independent of $N.$ According to Degli Esposti et al \cite{degli}, this implies that the quantum map is ergodic and mixing in the semiclassical limit, within the definition of Von Neumann \cite {vonneum}. The average behavior of the QPF was established by Keating \cite {keat2} and it is shown that in the semiclassical limit ($N\rightarrow \infty $) \begin{equation} \left\langle \log \frac{k(N)}N\right\rangle \approx \frac{-2\sqrt{\pi }}{e^2}% \left( \log \log N\right) \left( \log \log \log N\right) \end{equation} that is, the degeneracy grows very slowly with $N.$ \begin{figure}[th] \centerline {\epsfxsize=6in \epsfysize=4in \epsffile{fig2.ps} } \caption{ The QPF for the Hannay and Berry cat map, chaotic with one degree of freedom, whose symplectic matrix is ${\cal M}_{hb},$ defined in (\ref{mhb}). } \label{fig.2} \end{figure} \subsubsection*{ The double hyperbolic case} We now choose the symmetric matrix $\beta _{hh}\,$ whose associated symplectic matrix is given by ${\cal M}_{hh}$ defined in (\ref{mhh}). The QPF of this map is shown in figure ~\ref{fig.3}. As we can see, the behavior is very different of that obtained for the Hannay and Berry Cat map. There are now families of parabolas instead of straight lines. The role played by $N$ in figure ~\ref{fig.2} is here played by $N^2$, because there are here $N^2$ states for this system. We conjecture that this kind of behavior implies quantal ergodicity and mixing for two degrees of freedom systems in the semiclassical limit, although a more formal study of this conjecture will be realized in a future work. \begin{figure}[th] \centerline {\epsfxsize=6in \epsfysize=4in \epsffile{fig3.ps} } \caption{ The QPF for the double hyperbolic map, whose symplectic matrix is $% {\cal M}_{hh},$ defined in ($\ref{mhh}$). } \label{fig.3} \end{figure} \subsubsection*{ The mixed case} We first consider the map whose symmetric matrix is $\beta _{eh1},$ defined in (\ref{meh1}), whose QPF is shown in figure ~\ref{fig.4}. This is compared to $\beta _{eh2}$, defined in (\ref{meh2}), characterized by irrational rotation angles, whose QPF is now shown in figure ~\ref{fig.5}. There is a marked difference in the behavior between figure ~\ref{fig.4}. and figure ~\ref {fig.5}. While the former is very similar to the one obtained in figure ~\ref {fig.2} for a chaotic one degree of freedom system, the one of figure ~\ref {fig.5}. is close to figure ~\ref{fig.3}. for a system of two degrees of freedom. The very different behavior, as was explained in section 3, is due to the fact that the classical eigenvalues $\lambda _3^{eh1}$ and $\lambda _4^{eh1}$ denote rotation by angles that are a fraction of $\pi .$ Then the behavior of the system will be equivalent a hyperbolic system with a single degree of freedom and it is ergodic only in a subspace. On the contrary, $% \lambda _3^{eh2}$ and $\lambda _4^{eh2}$ denote rotation angles that are irrational fractions of $\pi ,$ so that classical and semiclassically the ergodicity appears in the whole phase space. \begin{figure}[tbp] \centerline {\epsfxsize=6in \epsfysize=3.5in \epsffile{fig4.ps} } \caption{ The QPF for the mixed map , with rational rotations, whose symplectic matrix is ${\cal M}_{eh1},$ defined in (\ref{meh1}). } \label{fig.4} \end{figure} \begin{figure}[tbp] \centerline {\epsfxsize=6in \epsfysize=3.5in \epsffile{fig5.ps} } \caption{ The QPF for the mixed map, with irrational rotations, whose symplectic matrix is ${\cal M}_{eh2},$ defined in (\ref{meh2}) .} \label{fig.5} \end{figure} \clearpage \subsubsection*{ The loxodromic case} Let us now study the map denoted by the $\beta _{lox}$ matrix defined by (% \ref{mlox}). The corresponding QPF is shown in figure ~\ref{fig.6}. The behavior is then similar to the one obtained for ergodic systems with two degrees of freedom. This map then manifests classical and semiclassical ergodicity on the whole phase space or the corresponding Hilbert space. Similarly to the previous example, we found that rotation angles that are a fraction of $\pi $ behave similarly to hyperbolic one degree of freedom maps shown in figure ~\ref{fig.5}. \begin{figure}[h] \centerline {\epsfxsize=6in \epsfysize=3.5in \epsffile{fig6.ps} } \caption{ The QPF for the double loxodromic map of two degrees of freedom, with irrational rotations, whose symplectic matrix is ${\cal M}_{lox},$ defined in (% \ref{mlox}).} \label{fig.6} \end{figure} \clearpage \section{The quantum feline group} In this section we present the general quantization of multidimensional cat maps constructed in section two. In a first step we establish the conditions for the maps to be quantizable and then we construct their center and chord representations, based on the formalism developed in reference \cite{opetor} and summarized in the Appendix. The classical automorphism generated by ${\cal M}\,$ in plane phase space being linear, its quantization on the Hilbert space $\left[ {\cal {H}_{{\Bbb R}}}% \right] ^L$ ,associated to the Euclidean phase space, will have the crucial property that \begin{equation} \widehat{U}_{\cal M}\widehat{T}_\xi \widehat{U}_{\cal M}^{\dagger }=\widehat{T}_{{\cal M}\xi } \label{utu} \end{equation} for any translation operator $\widehat{T}_\xi $. We now show how to associate to any $\widehat{U}_{\cal M}$ a unitary operator on $\left[ {\cal H}% _N^\chi \right] ^L$, the torus Hilbert space characterized by the Floquet parameter $\chi .$ For this purpose let us study the restriction of $% \widehat{U}_{\cal M}$ to $\left[ {\cal H}_N^\chi \right] ^L.$ We then have: {\bf Proposition} \begin{equation} \widehat{U}_{\cal M}\left[ {\cal H}_N^\chi \right] ^L\subset \left[ {\cal H}% _N^{\chi ^{\prime }}\right] ^L \end{equation} {\it where } \begin{equation} \chi ^{\prime }={\cal M}\chi -\frac N2{\frak J}\left( {\cal M}\otimes {\cal M}\right) \quad % \mbox{mod(1)} \label{xixip} \end{equation} {\it where we have defined the vector product of two matrices as the vector whose components are } \begin{equation} \left( A\otimes B\right) _i=\sum_{j=1}^LA_{i,j}B_{i,j+L} \end{equation} {\bf Proof: }Equation (\ref{utu}) implies \begin{equation} \widehat{U}_{\cal M}\widehat{T}_\xi =\widehat{T}_{{\cal M}\xi }\widehat{U}_{\cal M}. \label{tuut} \end{equation} We now restrict considerations to chords that perform integer loops around the torus, $\xi ={\bf m,}$ with ${\bf m\,\,}$ an integer vector with $2L$ components,i.e., chords that generate translations that are equivalent to the identity on the torus. We thus act on a given state of $\left[ {\cal H}% _N^\chi \right] ^L$ on each side of (\ref{tuut}), so that inserting (\ref {tNQ}), we obtain \begin{equation} e^{\left[ i2\pi N\left( \frac 14{\bf m}\widetilde{{\frak J}}{\bf m}+{\bf m}% \wedge \frac \chi N\right) \right] }\widehat{U}_{\cal M}|{\bf \Psi }>=e^{\left[ i2\pi N\left( \frac 14{\bf m}{\cal M}^t\widetilde{{\frak J}}{\cal M}{\bf m}+{\cal M}{\bf m}\wedge \frac{\chi ^{\prime }}N\right) \right] }\widehat{U}_{\cal M}|{\bf \Psi }>, \end{equation} for any vector ${\bf m}$. Now we chose ${\bf n}={\cal M}{\bf m}$ , hence, we find that \begin{equation} {\bf n}\wedge \chi ^{\prime }={\cal M}^{-1}{\bf n}\wedge \chi +\frac N4{\bf n}% \left[ \left( {\cal M}^{-1}\right) ^t\widetilde{{\frak J}}\left( {\cal M}^{-1}\right) -% \widetilde{{\frak J}}\right] {\bf n}\qquad \mbox{mod}(1) \label{xixipk} \end{equation} and we specify $2L\,$independent integer vectors ${\bf n}\,=e_j% \,(j=1,...,2L),$ such that each of them denotes one loop around one of the $j $-irreducible circuit on the torus. In this case $e_j\widetilde{{\frak J}}% e_j=0,$ using the symplectic property of ${\cal M}$ (\ref{msimple}), and the fact that ${\frak J}\widetilde{{\frak J}}{\frak J}=\widetilde{{\frak J}}$, we obtain \begin{equation} e_j\wedge \chi ^{\prime }={\cal M}^{-1}e_j\wedge \chi +\frac N4e_j{\frak J}{\cal M}% \widetilde{{\frak J}}{\cal M}^t{\frak J}e_j\mbox{mod(1)}. \end{equation} Remarking that \begin{equation} \frac N4\left( {\frak J}{\cal M}\widetilde{{\frak J}}{\cal M}^t{\frak J}\right) _i=-\frac N2\sum_{j=1}^L{\cal M}_{i,j}{\cal M}_{i,j+L}, \end{equation} we finally obtain (\ref{xixip}) after straightforward manipulations. In the case $L=1$, (\ref{xixip}) reduces to the already known result \cite {degli},\cite{debievre}, \begin{equation} \chi ^{\prime }={\cal M}\chi -\frac N2{\frak J}\left( \begin{array}{l} {\cal M}_{11}{\cal M}_{12} \\ {\cal M}_{21}{\cal M}_{22} \end{array} \right) \quad \mbox{mod(1)}. \end{equation} For two degrees of freedom ($L=2$) we obtain: \begin{equation} \chi ^{\prime }={\cal M}\chi -\frac N2{\frak J}\left( \begin{array}{l} {\cal M}_{11}{\cal M}_{13}+{\cal M}_{12}{\cal M}_{14} \\ {\cal M}_{21}{\cal M}_{23}+{\cal M}_{11}{\cal M}_{24} \\ {\cal M}_{31}{\cal M}_{33}+{\cal M}_{32}{\cal M}_{34} \\ {\cal M}_{41}{\cal M}_{43}+{\cal M}_{42}{\cal M}_{44} \end{array} \right) \quad \mbox{mod(1)}. \label{xixip4} \end{equation} Given ${\cal M},$ there exists for each $N$ a set of solutions of (\ref{xixip}) such that $\chi ^{\prime }=\chi $. This Floquet set of parameters is determined by \begin{equation} \chi =\frac N2\left( {\cal M}-1\right) ^{-1}{\bf i}_{\cal M}+\left( {\cal M}-1\right) ^{-1}{\bf m,% } \label{xiquanti} \end{equation} where ${\bf i}_{\cal M}={\frak J}\left( {\cal M}\otimes {\cal M}\right) \mbox{mod(4)}\,$ is an integer vector. For this set of parameters the propagator $\widehat{U}_{\cal M}$ is such that \begin{equation} \left[ \widehat{U}_{\cal M},\widehat{{\bf 1}}_N^\chi \right] =0, \label{[u,1]} \end{equation} where $\widehat{{\bf 1}}_N^\chi $ is the unit operator in $\left[ {\cal H}% _N^\chi \right] ^L$ defined in (\ref{eq:1toro})$.$ Evidently we cannot change the Floquet parameters at each iteration of the map, so (\ref {xiquanti}) defines the {\it quantizability set of the map}. For this set of parameters, we can then restrict the unitary operators $\widehat{U}_{\cal M}\,$ to the Hilbert space of the $2L$-torus $\left[ {\cal H}_N^\chi \right] ^L$ and thus define the corresponding torus propagator as \begin{equation} \widehat{{\bf U}}_{\cal M}^\chi =\widehat{U}_{\cal M}\widehat{{\bf 1}}_N^\chi =\widehat{% {\bf 1}}_N^\chi \widehat{U}_{\cal M}\widehat{{\bf 1}}_N^\chi . \label{ugato} \end{equation} Operators defined through (\ref{ugato}) inherit the unitary from their plane counterpart and we will call them {\it quantum cat maps. } The fact that the quantum symplectic map commutes with projections on the torus (\ref{eq:abpt}) leads to; \begin{equation} \widehat{{\bf U}}_{{\cal M}^l}^\chi =\left[ \widehat{{\bf U}}_{\cal M}^\chi \right] ^l. \label{uml} \end{equation} The unitarity of $\widehat{{\bf U}}_{\cal M}^\chi $ and (\ref{uml}) implies that \begin{equation} \widehat{{\bf U}}_{\cal M}^\chi =\left[ \widehat{{\bf U}}_{\cal M}^\chi \right] ^{-1}=\left[ \widehat{{\bf U}}_{\cal M}^\chi \right] ^{\dagger }. \end{equation} Having two cat maps ${\cal M}_A$ and ${\cal M}_B$ that are quantizable on the same Floquet parameter $\chi ,$ implies that (\ref{eq:abpt}) \begin{equation} \widehat{{\bf U}}_{{\cal M}_A}^\chi \widehat{{\bf U}}_{{\cal M}_B}^\chi =\widehat{{\bf U}}% _{{\cal M}_A{\cal M}_B}^\chi , \label{ucompoL} \end{equation} that is, the quantization of the composition of different cat maps is equivalent to the composition of the quantization of each map. This shows that quantum cat maps form a group, the quantum propagator preserving the classical composition laws. This{\it \ feline quantum group} is at the same time a projection onto the torus of the metaplectic group and a subgroup of the classical feline group. It then follows that, for any power $l$, the map ${\cal M}^l$ is also quantizable. The quantizability set of the map can be made independent of $N$ if the inhomogeneous term in (\ref{xixip})\ vanishes. This will be the case for matrices ${\cal M}$, such that \begin{equation} {\bf m}\left( {\cal M}^t\widetilde{{\frak J}}{\cal M}\pm \widetilde{{\frak J}}\right) {\bf % m=}0\qquad \mbox{mod}(2) \label{Mquanti} \end{equation} for any integer vector ${\bf m.}$ This condition is equivalent to \begin{equation} \sum_{j=1}^L{\cal M}_{i,j}{\cal M}_{i,j+L}=\mbox{even}, \label{Mxi0} \end{equation} so that the allowed Floquet parameters are \begin{equation} \chi =\left( {\cal M}-1\right) ^{-1}{\bf m.} \end{equation} In conclusion, the Floquet parameters are then denoted by the fixed points of the map (\ref{xfix}). The QPS can thus have any of the fixed points of the map as its origin; without loss of generality, one can take $\chi =0.$ For the one degree of freedom case (\ref{Mxi0}) implies that the ${\cal M}$ matrix must be restricted to the family of maps \[ \left( \begin{array}{cc} \mbox{even} & \mbox{odd} \\ \mbox{odd} & \mbox{even} \end{array} \right) \qquad \mbox{or\qquad }\left( \begin{array}{cc} \mbox{odd} & \mbox{even} \\ \mbox{even} & \mbox{odd} \end{array} \right) \] selected as quantizable by Hannay and Berry \cite{hanay}. Let us now provide an explicit construction of quantum cat maps based on the center and chord representations of operators. We start with the quantization of the linear automorphism ${\cal M}$ on the Hilbert space $\left[ {\cal {H}_{{\Bbb R}}}\right] ^L$ associated to the Euclidean phase space. The linearity of the ${\cal M}$ map implies in the exactness of the Van Vleck construction of the propagator \cite{van vleck}. This propagator in the chord representation has the form \cite{ozrep} \begin{equation} U_{\cal M}(\xi )=\left| \det \left[ 1\pm {\frak J}\frac{\partial ^2S(\xi )}{% \partial \xi ^2}\right] \right| ^{\frac 12}e^{-\frac i\hbar S(\xi )} \label{upcor} \end{equation} where $S(\xi )\equiv S(\xi ,0)$ is the chord generating function of the automorphism in the plane phase space. In the center representation, the Van Vleck propagator is \begin{equation} U_{\cal M}(x)=\left| \det \left[ 1\pm {\frak J}\frac{\partial ^2S(x)}{\partial x^2}% \right] \right| ^{\frac 12}e^{\frac i\hbar S(x)}, \label{upx} \end{equation} where now $S(x)\equiv S(x,0)$ is the center generating function of the transformation in the plane. Inserting (\ref{scor}) and (\ref{sx}) for the respective generating functions, we have \cite{ozrep} \begin{equation} U_{\cal M}(x)=\left| \det \left( 1\pm {\frak J}B\right) \right| ^{-\frac 12}e^{\frac i\hbar xBx} \label{ugatox} \end{equation} and \begin{equation} U_{\cal M}(\xi )=\left| \det \left( 1\pm {\frak J}\beta \right) \right| ^{-\frac 12}e^{-\frac i\hbar \frac 14\xi \beta \xi }. \label{ugatocor} \end{equation} We may now project these representations on the $2L$-torus. For this purpose we must restrict our construction to the quantizability set of Floquet parameter $\chi $ defined through (\ref{xiquanti}) and take the chord and center symbols of the projected operator. This is performed substituting (% \ref{ugatocor}) and (\ref{ugatox}) respectively in (\ref{eq:Acorprom}) and (% \ref{eq:Axplator}). Then we obtain the chord representation of the quantum cat map as an average over winding vectors: \begin{equation} {\bf U}_{\cal M}^\chi (\xi )=\left| \det \left( 1\pm {\frak J}\beta \right) \right| ^{-\frac 12}\left\langle e^{-i2\pi N\left[ \frac 14\xi \beta \xi +\frac 12\xi (\beta +{\frak J}){\bf m}+\frac 14{\bf m}(\beta -\widetilde{{\frak J}})% {\bf m+}\frac \chi N\wedge {\bf m}\right] }\right\rangle _{{\bf m}}. \label{ugatocor2} \end{equation} There are various classical objects present in this formula. Recalling section 2, $S(\xi ,{\bf m}),$ the chord generating function of the cat map defined in (\ref{scor}) appears in the exponent of (\ref{ugatocor2}). In the amplitude we recognize $\tau _\xi ,$ the number of fixed points of the map defined in (\ref{d1}), so that, \begin{equation} {\bf U}_{\cal M}^\chi (\xi )=2^L\frac 1{\sqrt{\tau _\xi }}\left\langle e^{-i2\pi N\left[ S(\xi ,{\bf m}){\bf +}\frac \chi N\wedge {\bf m}\right] }\right\rangle _{{\bf m}}. \label{ugatocors} \end{equation} Similarly, for the center representation we obtain \begin{equation} {\bf U}_{\cal M}^\chi (x)=\left| \det \left( 1\pm {\frak J}B\right) \right| ^{-\frac 12}\left\langle e^{i2\pi N\left[ xBx+x(B-{\frak J}){\bf m}+\frac 14% {\bf m}(B+\widetilde{{\frak J}}){\bf m-}\frac \chi N\wedge {\bf m}\right] }\right\rangle _{{\bf m}}, \label{ugatox2} \end{equation} where we recognize $S(x,{\bf m}),$ the center generating function of the cat map (\ref{sx}), and $\tau _x,$ the number of orbits centered on $x$ defined by (\ref{gama}), so that \begin{equation} {\bf U}_{\cal M}^\chi (x)=2^L\frac 1{\sqrt{\tau _x}}\left\langle e^{i2\pi N\left[ S(x,{\bf m}){\bf -}\frac \chi N\wedge {\bf m}\right] }\right\rangle _{{\bf m}% }. \label{ugatoxs} \end{equation} Both (\ref{ugatocors}) and (\ref{ugatoxs}) are representations of quantum cat maps of general dimension, showing that the quantum propagator is entirely defined in terms of classical objects, except for the term $\frac \chi N\wedge {\bf m}$ in the exponent that denotes the quantum features of the boundary conditions. However, we have to perform the average on ${\bf m}$ to make the representation explicit. Given that the quantizability set (\ref {xiquanti}) only admits rational values of $\chi $, (\ref{ugatocors}) and (% \ref{ugatoxs}) are Gaussian sums. The quantizability condition (\ref{Mquanti}% ) for the map ${\cal M}$ is equivalent to the condition that the Gaussian sums do not vanish for $L=1.$ We do not know of a similar verification for $L>1.$ In the following we will restrict our attention to maps that fulfill the condition (\ref{Mquanti}), such that the inhomogeneous term in (\ref{xixip}) vanishes. As we have already discussed, we can then choose $\chi =0$ without loss of generality. The quantum cat maps associated with these Floquet parameters will be denoted by $\widehat{{\bf U}}_{\cal M}$. All the examples in section 4 are of this type. We now use the periodicity properties (\ref{scorper}) of $S(\xi ,{\bf m})$ to show that the exponential function in (\ref{ugatocors}) is periodic. Indeed, for ${\bf m}^{\prime }$ defined in (\ref{mprim}) and the generating function $S(\xi ,{\bf m}^{\prime })$ defined in (\ref{scorper}) we have that \begin{equation} e^{-i2\pi N\left[ S(\xi ,{\bf m}^{\prime })\right] }=e^{-i2\pi N\left[ S(\xi ,{\bf m})\right] }e^{-i2\pi N\left[ \xi \wedge {\bf k-}\frac 12{\bf m}\Gamma _1{\bf k-}\frac 14{\bf k}\Delta _1{\bf k}\right] }, \end{equation} where $\Gamma _1$ and $\Delta _1$ were respectively defined in (\ref{gama1}) and (\ref{delta1}). We can now see that $e^{-i2\pi N\left[ \xi \wedge {\bf k}% \right] }=1$ for any integer vector ${\bf k}$ and for maps that fulfill the condition (\ref{Mquanti}) $e^{-i2\pi N\left[ {\bf -}\frac 12{\bf m}\Gamma _1% {\bf k-}\frac 14{\bf k}\Delta _1{\bf k}\right] }=1,$ hence \begin{equation} \exp \left[ -i2\pi NS\left( \xi ,{\bf m}+({\cal M}-1){\bf k}\right) \right] =\exp \left[ -i2\pi NS(\xi ,{\bf m})\right] . \end{equation} The average in (\ref{ugatocors}) is then periodic, so that, we sum over one period and divide by the number of points in the period, to perform such an average. Hence, we must restrict ${\bf m}$ to the fundamental parallelogram $% \Diamond _\xi $ defined in (\ref{para1}), where there are exactly $\tau _\xi $ points ${\bf m}$ with integer coordinates: \begin{eqnarray} {\bf U}_{\cal M}(\xi ) &=&2^L\left( \tau _\xi \right) ^{-\frac 32}\sum_{{\bf m\in \Diamond }_\xi }e^{-i2\pi N\left[ S(\xi ,{\bf m})\right] } \nonumber \\ &=&2^L\left( \tau _\xi \right) ^{-\frac 32}\sum_{{\bf m\in \Diamond }_\xi }e^{-i2\pi N\left[ \frac 14\xi \beta \xi +\frac 12\xi (\beta +{\frak J}){\bf % m}+\frac 14{\bf m}(\beta -\widetilde{{\frak J}}){\bf m}\right] }. \label{ugcorp} \end{eqnarray} Since ${\bf m}$ belongs to $\Diamond _\xi ,$ the sum in (\ref{ugcorp}) is the sum over the different classical orbits whose chord is $\xi $. In a similar way, for the center representation the periodicity region is the parallelogram $\Diamond _x\,,$ defined in (\ref{para2}), that has exactly $\tau _x$ points ${\bf m}$ with integer coordinates (\ref{gama}), so that \begin{eqnarray} {\bf U}_{\cal M}(x) &=&2^L\tau _x^{-\frac 32}\sum_{{\bf m\in \Diamond }_x}e^{i2\pi N\left[ S(x,{\bf m})\right] } \nonumber \\ &=&2^L\tau _x^{-\frac 32}\sum_{{\bf m\in \Diamond }_x}e^{i2\pi N\left[ xBx+x(B-{\frak J}){\bf m}+\frac 14{\bf m}(B+\widetilde{{\frak J}}){\bf m}% \right] }, \label{ugxp} \end{eqnarray} i.e. we are taking the sum on the different classical orbits centered in $x.$ The expressions (\ref{ugcorp}) and (\ref{ugxp}) have exactly the form expected for the semiclassical approximation of the propagator in center and chord representations respectively \cite{ozrep},\cite{opetor} but in this case they are exact instead of being a mere approximation. For generic cat maps, (\ref{ugcorp}) and (\ref{ugxp}) generate Gaussian sums, but in some cases important simplifications are possible. Let us start, once more, with the chord representation. We have already seen that the matrix $\beta $ has the form $\beta =\frac{\overline{\beta }^{\prime }}{% \tau _\xi ^{\prime }}$, where the barred matrix has integer elements$.$ The simplest case is when the cat map is such that the associated $\beta \,$% matrix itself has integer entries. Recalling that $\xi =\frac 1N\overline{% \xi }$, we will transform to an equivalent set of chords, for which the values for $\overline{\xi }$ are even multiples of $\tau _\xi ^{\prime }$. So, for any chord $\xi $ there is an equivalent chord $\Xi ,$ such that \begin{equation} \Xi =\xi +{\bf n}=\frac{2\tau _\xi ^{\prime }}N\overline{\Xi }, \label{corN} \end{equation} where the components of $\overline{\Xi }$ are integer numbers up to $N$ and the components of ${\bf n}$ are integer up to $2\tau _\xi ^{\prime }-1$. Equation (\ref{corN}) has solutions with the specified features for any $\xi ,$ only if $N\,$ and $2\tau _\xi ^{\prime }\,$ are coprime numbers. We will then restrict $N$ to be an odd integer. In this case, the chords $\Xi $ in (% \ref{corN}) form a lattice with spacing $\frac{2\tau _\xi ^{\prime }}N.$ A hypercube of side $2\tau _\xi ^{\prime }$ has then $N^L\times N^L$ successive chords $\Xi $ that constitute a basis for translation operators. For the simplest case where the matrix $\beta $ has integer entries, the chords $\Xi \,$ form a lattice of length $2$ with spacing $\frac 2N.$ Performing the transformation to chords $\Xi ,$\ we see that the term $% e^{-i2\pi N\left[ \frac 12\Xi (\beta +{\frak J}){\bf m}\right] }=1,$ so there is then no $\Xi $-dependence in the propagator sum (\ref{ugcorp}), leading to(\ref{ugcorimp}). In the same way, the matrix $B$ in the center representation has the form $B=% \frac{\overline{B}^{\prime }}{\tau _x^{\prime }}$ . Then we transform to center points $X$ that are integer multiples of $\frac{\tau _x^{\prime }}N$, \begin{equation} X=x+\frac 12{\bf j}=\frac{\tau _x^{\prime }}N\overline{X}, \label{xN} \end{equation} where the components of $\overline{X}$ are integer numbers up to $N$ and the components of ${\bf j}$ are integer up to $2\tau _x^{\prime }-1$. Again, solutions of (\ref{xN}) with the specified features will exist only if $N$ and $2\tau _x^{\prime }$ are coprime numbers. Thus, reflection operators on points $X$, that form a lattice with separation $\frac{\tau _x^{\prime }}N$ on a hypercube of side $\tau _x^{\prime },$ form a basis for the Hilbert space of the torus. For center points $X$ in (\ref{xN}) , the term $e^{i2\pi N\left[ X(B-{\frak J}){\bf m}\right] }=1,$ so the propagator (\ref{ugxp}) has the form (\ref{ugxd}). The above are then special cases where the propagator on the torus has the same form as its equivalent on the plane. These cases are then ideal to study quantization, as we have seen in section 4. We will now discuss the quantum effect of a similarity transformation of the form \begin{equation} {\cal M}\rightarrow {\cal M}^{\prime }={\cal N}^{-1}{\cal M}{\cal N}. \end{equation} As we have already discussed the matrices ${\cal M}^{\prime }$ and ${\cal M}$ represent the same map, but seen on a different frame of canonical coordinates. On the torus, we have to restrict both ${\cal M}$ and ${\cal N}$ to have integer elements so that the torus is mapped on itself. This restricts the symplectic similarity transformations to the feline transformations. The main advantage of the chord and center representation in plane phase space is their symplectic invariance\cite{ozrep}. It is well known that linear classical canonical transformations $x^{\prime }={\cal N}x$ correspond to unitary transformations in $\left[ {\cal {H}_{{\Bbb R}}}\right] ^L$, \begin{equation} \widehat{A}\rightarrow \widehat{A}^{\prime }=\widehat{U}_{\cal N}\widehat{A}% \widehat{U}_{\cal N}^{-1}. \end{equation} The effect of such a unitary transformation on the chord and center representation is merely \begin{equation} A(x)\rightarrow A({\cal N}x)\quad \mbox{and}\quad A(\xi )\rightarrow A({\cal N}\xi ). \end{equation} Because of the commutation of operator products with projection from the plane to the torus, the effect of a similarity transformation $\widehat{{\bf % A}}\rightarrow \widehat{{\bf U}}_{\cal N}^\chi \widehat{{\bf A}}\left[ \widehat{% {\bf U}}_{\cal N}^\chi \right] ^{-1}$ performed by a quantized cat map on any operator $\widehat{{\bf A}}$ that commutes with $\widehat{{\bf 1}}_{\cal N}^\chi $ will be purely classical in the center or the chord representations: \begin{equation} {\bf A}(x)\rightarrow {\bf A}({\cal N}x)\quad \mbox{and\quad }{\bf A}(\xi )\rightarrow {\bf A}({\cal N}\xi ). \label{cat1} \end{equation} Thus the similarity transformation among quantum cat maps, reduces to the classical similarity transformation: \begin{equation} \widehat{{\bf U}}_{{\cal M}^{\prime }}^\chi =\widehat{{\bf U}}_{\cal N}^\chi \widehat{{\bf % U}}_{\cal M}^\chi \left[ \widehat{{\bf U}}_{\cal N}^\chi \right] ^{-1}, \label{uprim} \end{equation} so that, \begin{equation} {\bf U}_{{\cal M}^{\prime }}(x)={\bf U}({\cal N}x)\quad \mbox{and\quad }{\bf U}_{{\cal M}^{\prime }}(\xi )={\bf U}({\cal N}\xi ). \label{catcat} \end{equation} However, the quantum cat maps ${\cal M}^{\prime }$ and ${\cal M}$ must be quantized on the same Floquet parameters $\chi $. This imposes another restriction on the matrix ${\cal N}$ used to perform the similarity transformation, its quantizability set must include the Floquet parameter $\chi .$ That is ${\cal M}$ and ${\cal N}$ must belong the same quantum feline group. For two quantum cat maps $\widehat{{\bf U}}_{{\cal M}^{\prime }}^\chi $ and $% \widehat{{\bf U}}_{\cal M}^\chi $ related by (\ref{uprim}) and for all power $l$ of the map we have \begin{eqnarray} Tr\left[ \left( \widehat{{\bf U}}_{{\cal M}^{\prime }}^\chi \right) ^l\right] &=&Tr\left\{ \left( \widehat{{\bf U}}_{\cal N}^\chi \widehat{{\bf U}}_{\cal M}^\chi \left[ \widehat{{\bf U}}_{\cal N}^\chi \right] ^{-1}\right) ^l\right\} =Tr\left\{ \left( \widehat{{\bf U}}_{\cal M}^\chi \right) ^l\left[ \widehat{{\bf U}}_{\cal N}^\chi \right] ^{-l}\left( \widehat{{\bf U}}_{\cal N}^\chi \right) ^l\right\} \nonumber \\ &=&Tr\left[ \left( \widehat{{\bf U}}_{\cal M}^\chi \right) ^l\right] . \label{TRML} \end{eqnarray} We have seen in (\ref{romapa}) that the density of states and hence the spectrum of the system is uniquely determined through the traces of the different powers of the map. Then, (\ref{TRML}) shows that $\widehat{{\bf U}}% _{{\cal M}^{\prime }}^\chi $ and $\widehat{{\bf U}}_{\cal M}^\chi $ related by (\ref{uprim}% ) have the same quasi-energy spectrum. We now show that the quantum period function is also invariant with respect to a feline similarity transformation, for the periodic case where we can chose the Floquet parameter $\chi =0$. Suppose that the QPF corresponding to $\widehat{{\bf U}}_{{\cal M}^{\prime }}^\chi $ and $\widehat{{\bf U}}_{\cal M}^\chi $ are respectively $k_{{\cal M}^{\prime }}(N)$ and $k_{\cal M}(N)$ , then \begin{equation} \left( {\cal M}^{\prime }\right) ^{k_{\cal M}(N)}=\left( {\cal N}^{-1}{\cal M}{\cal N}\right) ^{k_{\cal M}(N)}={\cal N}^{-1}\left( {\cal M}\right) ^{k_{\cal M}(N)}{\cal N}. \end{equation} For all points $x$ belonging to the QPS, i.e., a lattice of spacing $\frac 1N $ on the $2L$-torus $\Box ,$ we have \begin{equation} \left( {\cal M}^{\prime }\right) ^{k_{\cal M}(N)}x={\cal N}^{-1}\left( {\cal M}\right) ^{k_{\cal M}(N)}{\cal N}x=x, \end{equation} hence \begin{equation} \left( {\cal M}^{\prime }\right) ^{k_{\cal M}(N)}={\bf 1\quad }\mbox{mod}(N), \end{equation} so that \begin{equation} k_{{\cal M}^{\prime }}(N)\leq k_{\cal M}(N). \end{equation} In the same way, for all points $x$ belonging to the QPS \begin{equation} \left( {\cal M}\right) ^{k_{{\cal M}^{\prime }}(N)}x={\cal N}\left( {\cal M}^{\prime }\right) ^{k_{{\cal M}^{\prime }}(N)}{\cal N}^{-1}x=x, \end{equation} hence \begin{equation} k_{\cal M}(N)\leq k_{{\cal M}^{\prime }}(N). \end{equation} Therefore, \begin{equation} k_{{\cal M}^{\prime }}(N)=k_{\cal M}(N), \end{equation} $\,$ that is, the QPF of both quantum maps $\widehat{{\bf U}}_{{\cal M}^{\prime }}^\chi $ and $\widehat{{\bf U}}_{\cal M}^\chi $ coincide. \section{Conclusions} \setcounter{equation}{0} In this work we have studied classical and quantum properties of multidimensional Cat maps. In a first step the classical study was performed using the symplectically invariant center and chord generating functions. They allow us to represent the symplectic matrix by a symmetric one. This is the basis for a complete classification of generic four dimensional cat maps. Clearly, the advantage of working with the appropriate generating functions will be even more pronounced for cat maps of higher dimension. Two degrees of freedom are sufficient to obtain all distinct types of dynamics. Loxodromic behavior appears as a new alternative with respect to usual cat maps with one degree of freedom. The quantization of cat maps was performed using the recently developed Weyl representation and its conjugate chord representation. The semiclassical approximation is exact whatever the number of degrees of freedom or the characteristics of the cat map. The spectral properties show the same kind of ''pathologies '' observed for systems with one degree of freedom. Through the quantum periodicity function, we have indication of quantal ergodicity and mixing in the semiclassical limit for systems that present this classical property. We must note that this is one of the first times that loxodromic behavior is quantized \cite{pedrao}. According to Anosov's theorem, all cases of fully ergodic classical maps are structurally stable, that is, a weak nonlinear perturbation leads to a map whose orbits are topologically equivalent to the original cat map. The possibility of quantizing such an Anosov map is in no way restricted to one degree of freedom. In this way, one can obtain continuous families of quantum torus maps, corresponding to fully chaotic classical maps for each type of map ( doubly hyperbolic, loxodromic, etc.). These nonlinear maps will probably avoid the spectral anomalies due to quantum periodicity, as was verified for the case of a single degree of freedom \cite{matos} ,\cite{boas}. {\it Acknowledgments: }We thanks helpful discussions with J.P. Keating, who provide us Greenman preprint, E. Pujals and G. Contreras who provide us Ma\~ne's work. We acknowledge financial support from Pronex-MCT and A.M.F.R. also thanks support from CLAF-CNPq. This work was partially supported by contracts ANPCYT PICT97-01015, CONICET PIP98-420 and EC-931005AR.
\section{Introduction} The study of the weak radiative decay $B \to K^* \gamma $ as a test of Standard Model (SM) has attracted considerable attention since the CLEO experiment \cite{ref1} gave the preliminary determination of the exclusive branching ratio $Br (B \to K^* \gamma )=(4.0 \pm 1.7 \pm 0.8 ) \times 10^{-5} $. The weak radiative decays of $B$ mesons (which proceed through a flavor changing neutral current, absent at the tree level in the SM) are remarkable for several reasons. The $B \to K^* \gamma $ decay arises from the quark level process $b \to s \gamma $ via penguin-type diagrams at the one loop level. Hence it is not only a significant test of Standard Model flavor-changing neutral current dynamics but also is sensitive to new physics appearing through virtual particles such as the top quark and $W$ boson in the internal loop. The study of this process provides valuable information concerning the Cabibbo-Kobayashi-Maskawa (CKM) parameters, $V_{td},~V_{ts}$ and $V_{tb}$. Furthermore, additional contributions in loop stemming from new bosons and fermions present in most of the extensions of the SM, suggests the possibility that this process provides a window to new physics. The theoretical analysis of the decay process $B \to K^* \gamma $ requires long-distance QCD contributions which cannot be determined perturbatively. It is also not straightforward to calculate the exclusive decays by first principles of QCD, due to complications inherent in nonperturbative QCD. Therefore, one must resort to some phenomenological models to obtain reliable results. The heavy quark effective theory (HQET) \cite{ref2} is expected to be useful in this regard in so far as the $b$ quark is concerned. However the $s$ quark in the final hadron can neither be considered heavy enough to enable the use of HQET nor sufficiently light to permit the exploitation of the chiral perturbation theory in an unambiguous manner. There are several methods available in the literature to study the exclusive $B\to K^* \gamma $ decay process. Some of them include the QCD sum rule,\cite{ref3,ref31,ref32} lattice QCD,\cite{ref4} nonrelativistic and relativistic quark models.\cite{ref5,ref6,ref7,ref8,ref9} The HQET \cite{ref10} has also been applied to this decay process even though the $s$-quark mass is certainly not heavy enough, contrary to the requirement of HQET. In this paper we study the rare radiative decay process $B \to K^* \gamma $ using the covariant oscillator quark model (COQM).\cite{ref11} One of the most important motives for this model is to describe covariantly the center of mass motion of hadrons, while preserving the considerable success of the non-relativistic quark model regarding the static properties of hadrons. A keystone in COQM for doing this is treating directly the squared masses of hadrons in contrast to the mass itself, as done in conventional approaches. This makes the covariant treatment simple. The COQM has been applied to various problems \cite{ref12} with satisfactory results. Recently Ishida et al. \cite{ref13,ref181} have studied the weak decays of heavy hadrons using this model and derived the same relations of weak form factors for heavy-to-heavy transition as done in HQET.\cite{ref2} In addition, our model is also applicable to heavy-to-light transitions. As a consequence, this model does incorporate the features of heavy quark symmetry and can be used to compute the form factors for heavy-to-light transitions as well, which is beyond the scope of HQET. Actually, in previous papers we made analyses of the spectra of exclusive semi-leptonic\cite{ref181} decays of $B$-mesons and analyses of non-leptonic decays of $B$ mesons\cite{meson} and of hadronic weak decays of $\Lambda_b$ baryons\cite{baryon} using this line of reasoning, leading to encouraging results. Keeping this success in mind, we extend its application to weak radiative decays of $B$ mesons. The paper is organized as follows. In \S 2 we present the expressions for the decay widths for the weak radiative decays of $B$ mesons. In \S 3 we present a brief description of the covariant oscillator quark model. Using this model we have evaluated the required form factors. Section 4 contains our results and discussion. \section{Methodology} The general amplitude of weak radiative decay with one real photon emission is given by \begin{eqnarray} {\cal M}(B(p) &\to & P^*(k) \gamma(q) )=i \epsilon_{\mu \nu \alpha \beta}~\eta^\mu~q^\nu~ \epsilon^{\alpha}~k^{\beta}~ f_1(q^2) \nonumber\\ & +& \eta^{\mu}\left [\epsilon_{\mu}(M_B^2-M_{P^*}^2) -(p+k)_{\mu} (\epsilon \cdot q) \right ] f_2(q^2)\;, \label{eq1} \end{eqnarray} where $\eta $ and $\epsilon $ are the polarization vectors of the photon and the vector meson $P^* $, respectively. The first (second) term on the right-hand side of Eq. (\ref{eq1}) is a parity conserving (violating) term, and because of the real photon we have $q^2$=0. The decay width implied by this amplitude is given as \begin{equation} \Gamma(B \to P^* \gamma)= \frac{1}{32 \pi}\frac{\left (M_B^2 -M_{P^*}^2 \right )^3}{M_B^3} \left [|f_1(0)|^2 +4|f_2(0)|^2 \right ]\;. \label{eq2} \end{equation} Now, the exclusive decay $B \to K^* \gamma $ is expected to be described well by the quark level process $b \to s \gamma $ if the confinement effect is properly taken into account. Accordingly, in the Standard Model, $B$ decays are described by the effective Hamiltonian obtained by integrating out the top quark and $W$ boson fields given as \cite{ref14} \begin{equation} {\cal H}_{\rm eff}(b \to s \gamma) = -\frac{4G_F}{\sqrt{2}}\; V_{tb}\; V_{ts}^*\; \sum_{i=1}^8~C_i(\mu)~O_i(\mu)\;, \end{equation} where the $C_i(\mu)$ are the Wilson coefficients, arise from the renormalization group equation to provide the scaling down of the subtraction point appropriate to the problem i.e., $ \mu \approx m_b $. The set $\{O_i\} $ is a complete set of renormalized, dimension-six operators which govern the $b \to s$ transition. They consist of two current operators $O_1$ and $O_2$ and four strong penguin operators $O_3$--$O_6$, which determine the nonleptonic decays, and the electromagnetic dipole operator $O_7 $ and the chromomagnetic dipole operator $O_8$, which are responsible for rare $B$ decays, i.e., $b \to s+\gamma $ and $b \to s+g $, respectively. Out of the eight operators $O_i$, the one that contributes to $B \to K^* \gamma $ is \begin{equation} O_7=\frac{ e}{32 \pi^2} F_{\mu \nu}~ \left [ m_b \bar s ~\sigma^{\mu \nu}~(1+\gamma_5)~b +m_s \bar s ~\sigma^{\mu \nu}~(1-\gamma_5)~b \right ]\;, \end{equation} where $F_{\mu \nu} $ is the electromagnetic field strength tensor. The relevant Wilson coefficient is given by \cite{ref14} \begin{eqnarray} C_7(\mu) &=& \left [ \frac{\alpha_s(M_W)}{\alpha_s(m_b)} \right ]^{16/23} \biggr\{C_7(M_W)-\frac{8}{3}C_8(M_W)\nonumber\\ &\times &\left [1-\left (\frac{\alpha_s(m_b)}{\alpha_s(M_W)} \right )^{2/23} \right ] +\frac{232}{513} \left [1-\left (\frac{\alpha_s(m_b)}{\alpha_s(M_W)} \right )^{19/23} \right ] \biggr\}\;, \end{eqnarray} with \begin{equation} C_7(M_W)=-\frac{x}{2}\left [ \frac{\frac{2}{3}x^2+ \frac{5}{12}x-\frac{7}{12}}{(x-1)^3}- \frac{(\frac{3}{2}x^2-x){\rm ln}x}{(x-1)^4} \right ] \end{equation} and \begin{equation} C_8(M_W)=-\frac{x}{4}\left [ \frac{\frac{1}{2}x^2- \frac{5}{2}x-1}{(x-1)^3}+ \frac{3x{\rm ln}x}{(x-1)^4} \right ]\;, \end{equation} where $x=m_t^2/M_W^2 $. From the fact that $m_b \gg m_s $, only the term involving $m_b $ in the operator $O_7 $ need be retained. Thus the matrix element of interest becomes \begin{equation} \langle K^* \gamma | O_7 | B \rangle =\frac{i e}{16 \pi^2} ~q_{\mu} \eta_{\nu}~m_b~ \langle K^*|\bar s~ \sigma^{\mu \nu}~(1+\gamma_5)~b |B \rangle \;, \label{eq8} \end{equation} where $q_{\mu} $ is the four momentum of the photon and $\eta_{\mu} $ is its polarization vector. It should be noted that here the matrix element of a tensor current between hadronic states for which not much information is available is involved. However, in the framework of HQET the associated heavy quark spin symmetry enables one to express the matrix element of the tensor operator in terms of the vector and axial vector form factors that also occur in semileptonic decays and may be estimated in different phenomenological models. In the static limit of a heavy $b$ quark we may use the equation of motion $\gamma_0 b=b $ to derive the relations \cite{ref15} \begin{equation} \langle K^*|\bar s~i~\sigma_{0i}~(1+\gamma_5)~b|B \rangle =\langle K^*|\bar s~\gamma_{i}~(1-\gamma_5)~b|B \rangle \;. \label{eq:eqn1} \end{equation} As a result, the form factors $f_1 $ and $f_2$ in Eq. (\ref{eq1}) can be related to the vector and axial vector form factors $V$ and $A_1$ appearing in the matrix element of the RHS of Eq. (\ref{eq:eqn1}) defined by \cite{ref16} \begin{equation} \langle K^* (k) | \bar s \gamma_{\mu} b |B(p) \rangle =\frac{2i}{M_B+M_{K^*}} \epsilon_{\mu \nu \alpha \beta}~ \epsilon^{\nu} q^{\alpha}p^{\beta}~ V(q^2)\;,\label{eq:eqn3} \label{eq10} \end{equation} \begin{eqnarray} \langle K^* (k) | \bar s \gamma_{\mu}\gamma_5 b |B(p) \rangle &=&(M_B+M_{K^*}) \epsilon_{\mu}~A_1(q^2)-\frac{\epsilon \cdot q}{M_B+M_{K^*}}(p+k)_{\mu}~A_2(q^2)\nonumber\\ & & -2\frac{\epsilon \cdot q}{q^2} q_{\mu}~ M_{K^*}~[A_3(q^2) -A_0(q^2)]\;, \label{eq:eqn4} \end{eqnarray} with $q=p-k$. Here $A_3(q^2)$ is simply an abbreviation for \begin{equation} A_3(q^2)=\frac{M_B + M_{K^*}}{2 M_{K^*}}~A_1(q^2) -\frac{M_B-M_{K^*}}{2M_{K^*}}~A_2(q^2)\;, \label{eq12} \end{equation} and in order to cancel the singularity at $q^2=0$, we must have $A_3(q^2=0)=A_0(q^2=0)$. Thus with Eqs. (\ref{eq1}) and (\ref{eq8})--(\ref{eq12}), we obtain the following relations for $B \to K^* $ transition at $q^2=0 $:\cite{ref17} \begin{equation} f_1(0)=\frac{G_F}{\sqrt 2} \frac{ e}{2 \pi^2}~C_7(\mu)~ V_{tb} V_{ts}^*~ m_b ~F(0) \label{eq13} \end{equation} and \begin{equation} f_2(0)=\frac{1}{2}f_1(0), \label{eq14} \end{equation} where \begin{equation} F(0)=\frac{M_B -M_{K^*}}{2M_B}V(0)+\frac{M_B+M_{K^*}}{2M_B} A_1(0). \label{eq15} \end{equation} Substituting the above values of $f_1$ and $f_2$ into Eq. (\ref{eq2}), we obtain the decay width for $B \to K^* \gamma $ as \begin{eqnarray} \Gamma(B \to K^* \gamma)&=&\frac{\alpha~G_F^2~m_b^2} {128\pi^4 M_B^5} ~ |V_{ts} V_{tb}|^2~|C_7(\mu)|^2 (M_B^2 -M_{K^*}^2)^3 \nonumber\\ & & \times \biggr[ (M_B +M_{K^*})A_1(0)+(M_B-M_{K^*})V(0) \biggr]^2\;. \label{eq:eqn2} \end{eqnarray} Similarly, the decay width for the CKM suppressed FCNC radiative transition $b \to d \gamma $, which is responsible for $B^- \to \rho ^- \gamma $ and $B_s \to K^* \gamma $, is obtained from Eq. (\ref{eq:eqn2}) with $M_B $ and $M_{K^*} $ replaced by the corresponding initial and final meson masses. The relevant CKM factor in this case is $|V_{tb} V_{td}|^2 $. Now to evaluate the form factors $A_1(0) $ and $V(0) $, we use the covariant oscillator quark model, which is presented in the next section. Here it should be noted that the relations (\ref{eq13}) and (\ref{eq14}) are also derivable\cite{ref13} directly from Eq. (\ref{eq8}) in COQM, by using the covariant spin wave functions, that is, the Bargmann-Wigner spinor functions. Accordingly, the expression of $F(0)$ in terms of a space-time wave function (Eq. (\ref{eq26}) given later) is also derived directly. \section{ Model framework and the hadronic form factors} The general treatment of COQM may be called the ``boosted $LS$-coupling scheme,'' and the wave-functions, being tensors in $\tilde U(4) \times O(3,1) $-space, reduce to those in $SU(2)_{\rm spin} \times O(3)_{\rm orbit} $-space in the nonrelativistic quark model in the hadron rest frame. The spinor and space-time portion of the wave functions satisfy separately the respective covariant equations, the Bargmann-Wigner (BW) equation for the former, and the covariant oscillator equation for the latter. The form of the meson wave function has been determined completely through the analysis of mass spectra. In COQM the meson states are described by bi-local fields $\Phi_A^B(x_{1\mu},x_{2\mu}) $, where $x_{1\mu}(x_{2\mu})$ is the space time coordinate of the constituent quark (antiquark), $A=(a,\alpha)\;(B=(b,\beta))$, describing its flavor and covariant spinor. Here we write only the positive frequency part of the relevant ground state fields: \begin{equation} \Phi_A^B(x_{1\mu},x_{2\mu})=e^{iP\cdot X}\; U(P)_A^B\; f_{ab}(x_{\mu};P)\;, \end{equation} where $U$ and $f$ are the covariant spinor and internal space-time wave functions, respectively, satisfying the Bargmann-Wigner and oscillator wave equations. The quantity $x_{\mu} ~(X_{\mu}) $ is the relative (CM) coordinate, $x_{\mu}\equiv x_{1\mu}-x_{2\mu} ~(X_{\mu}\equiv m_1x_{1\mu} +m_2x_{2\mu})/(m_1+m_2) $, with $m_i$ the quark masses). The function $U$ is given by \begin{equation} U(P)=\frac{1}{2 \sqrt 2}\left [(-\gamma_5 P_s(v)+i\gamma_{\mu} V_{\mu}(v))(1+iv \cdot \gamma)\right ], \end{equation} where $P_s(V_\mu )$ represents the pseudoscalar (vector) meson field, and $v_{\mu} \equiv P_{\mu}/M~ (P_{\mu}(M)$ is the four momentum (mass) of the meson). The function $U$, being represented by the direct product of a quark and antiquark Dirac spinor with the meson velocity, is reduced to the non-relativistic Pauli-spin function in the meson rest frame. The function $f$ is given by \begin{equation} f(x_{\mu};P)=\frac{\beta}{\pi} \exp\left (-\frac{\beta}{2} \left (x_{\mu}^2+2\frac{(x\cdot P)^2}{M^2}\right ) \right )\;. \end{equation} The value of the parameter $\beta $ is determined from the mass spectra \cite{ref18} as $\beta_{\rho}=0.14 $, $\beta_{K^*}$ = 0.142, $\beta_B$ = 0.151 and $\beta_{B_s} $= 0.160 (in units of $ \mbox{GeV}^2$ ). The effective action for weak interactions of mesons with $W$-bosons is given by \begin{equation} S_W=\int d^4 x_1 d^4 x_2 \langle \bar \Phi_{F, P^\prime} (x_1,x_2)i \gamma_{\mu}(1+\gamma_5) \Phi_{I, P}(x_1,x_2) \rangle W_{\mu,q}(x_1)\;, \label{eq:eqn6} \end{equation} where we have denoted the interacting (spectator) quarks as 1 (2) (the KM matrix elements and the coupling constants are omitted). This is obtained from consideration of Lorentz covariance, assuming a quark current with the standard $V-A$ form. In Eq. (\ref{eq:eqn6}), $\Phi_{I,P}~ (\bar \Phi_{F,P^\prime})$ denotes the initial (final) meson with definite four momentum $P_\mu~(P_\mu^\prime) $, and $q_{\mu} $ is the momentum of $W$-boson. The function $\bar \Phi $ is the Pauli-conjugate of $\Phi $ defined by $\bar \Phi \equiv -\gamma_4 \Phi^\dagger \gamma_4 $, and $\langle ~~~\rangle$ represents the trace of Dirac spinor indices. Our relevant effective currents $J_\mu(X)_{P^\prime,P}$ are obtained\cite{ref13} by identifying the above action with \begin{equation} S_W =\int d^4 X J_{\mu}(X)_{P^\prime,P}\;W_{\mu}(X)_q, \end{equation} \begin{eqnarray} J_{\mu}&=&I^{qb}(w) \sqrt{M M^\prime} \times [\bar P_s(v^\prime)P_s(v)(v+v^\prime )_{\mu}\nonumber\\ & & +\bar V_{\lambda}(v^\prime ) P_s(v) (\epsilon_{\mu \lambda \alpha \beta} ~v_{\alpha}^{\prime}~v_\beta-\delta_{\lambda \mu} (w+1) - v_{\lambda}~v_{\mu}^{\prime}]\;, \label{eq:eqn5} \end{eqnarray} where $M (M^\prime) $ denotes the physical mass of the initial (final) meson and $w=-v\cdot v'$. The quantity $I^{qb}(w) $, the overlapping of the initial and final wave functions, represents the universal form factor function.\footnote{ In this paper we apply the pure-confining approximation, neglecting the effect of the one-gluon-exchange potential. This approximation is expected to be good for the heavy/light-quark meson systems. } It describes the confinement effects of quarks, and is given by \begin{equation} I^{qb}(w)=\frac{4 \beta \beta^\prime}{\beta+\beta^\prime } \frac{1}{\sqrt{C(w)}} \exp(-G(w))\;;~~ C(w)=(\beta-\beta^{\prime})^2+4\beta \beta^{\prime} w^2\;, \label{eq23} \end{equation} and \begin{eqnarray} G(w)&=& \frac{m_n^2}{2 C(w)} \biggr[ (\beta+\beta^\prime) \left \{ \left (\frac{M}{M_s} \right )^2 + \left (\frac{M^\prime} {M_s^\prime}\right )^2 -2 \frac{M M^\prime}{M_s M_s^\prime}w \right \}\nonumber\\ & &+ 2 \left \{ \beta^\prime\left (\frac{M}{M_s} \right )^2 +\beta\left ( \frac{M^\prime}{M_s^\prime}\right )^2 \right \}(w^2-1) \biggr]\;, \label{eq24} \end{eqnarray} where $M_s(M_s^\prime)$ represents the sum of the constituent quark masses of the initial (final) meson, and $m_n $ is the spectator quark mass. Comparing our effective current (\ref{eq:eqn5}) with Eqs. (\ref{eq:eqn3}) and (\ref{eq:eqn4}), we obtain\footnote{ Here, note the remark given at the end of \S 2. } the relations between the invariant form factors $V$ and $A_1$ with the form factor function $I(w)$ for $ B \to K^* $ transitions as \cite{ref181} \begin{eqnarray} V(0) &=& A_1(0)=F(0), \label{eq25}\\ F(0) &=& \frac{M_B+M_{K^*}}{2 \sqrt{M_B M_{K^*}}} I^{sb}(w|_{q^2=0})\;, \label{eq26} \end{eqnarray} and similarly for $B^- \to \rho^- $ and $B_s \to K^* $ transitions with $I^{db}(w) $ replacing $I^{sb}(w) $. \section{Results and conclusion} To estimate the branching ratios for weak radiative decays of $B$ mesons, we use the following values for various quantities. The quark masses (in GeV) are taken as $m_u=m_d$ = 0.4, $m_s$ = 0.55 and $m_b$ = 5. The particle masses and their lifetimes are taken from Ref. \citen{ref19}. The relevant CKM parameters required for all the processes considered here are taken from Ref. \citen{ref19} as $V_{td}=0.0085$, $V_{ts}= 0.0385$ and $V_{tb}=0.99925 $. The renormalized Wilson coefficient $C_7(m_b) $ used in the estimation of branching ratios is taken to be $|C_7(m_b)|$=0.311477.\cite{ref20} With the above values we first evaluate the form factor $F(0) $ using Eqs. (\ref{eq15}) and (\ref{eq23})--(\ref{eq25}). Our results for the form factor values given in Table I compare well with the recent calculations within the framework of the relativistic quark model,\cite{ref8} light cone QCD sum rule,\cite{ref31} and hybrid sum rule.\cite{ref32} With the calculated values of the form factors, the branching ratios for different channels are estimated using Eq. (\ref{eq:eqn2}) as well as the mean life values of the appropriate decaying meson, which are given in Table II. The branching ratios $Br(B^0 \to K^{*0} \gamma $ and $Br (B^{\pm} \to K^{*\pm} \gamma )$ are found to be in very good agreement with the available data. Here it may be worthwhile to note that the relevant process is relativistic, and the form factor function plays a significant role, such that $I=0.235$. Since there are no data as yet available for the branching ratios in the case $B \to \rho \gamma $ and $B_s \to K^* \gamma $, the model predictions must be compared with other theoretical predictions available in the literature. In fact there are few theoretical attempts made so far in this sector. Recently, Singer \cite{sing94} predicted that in the relativistic quark model, \begin{equation} Br(B_s \to K^* \gamma )=(1.4\pm 0.6) \times 10^{-6}, \end{equation} and in the relativistic independent quark model, Barik et al.\cite{ref9} obtained \begin{eqnarray} &&Br(B \to \rho \gamma)=1.24 \times 10^{-6},\nonumber\\ &&Br(B_s \to K^* \gamma)=9.28 \times 10^{-7}\;. \end{eqnarray} Finally, the hadronization ratio $R $ defined as $\Gamma(B \to K^* \gamma)/\Gamma(b \to s \gamma) $, where \begin{equation} \Gamma(b \to s \gamma) =\frac{G_F^2 \alpha}{32 \pi^4}~ ~|V_{ts} V_{tb}|^2~|C_7(m_b)|^2~ m_b^5\;, \end{equation} is found to be 0.12, which agrees well with the experimentally observed value $R=0.19 \pm 0.09.$ \cite{ref21} In view of the consistency of our predictions, obtained with no free parameters, with the large number of theoretical predictions as well as the experimental data, it may be concluded that the framework of COQM provides a suitable scheme to estimate the confined effect of quarks, and so far as the relevant flavor changing neutral current is concerned, the Standard Model is valid. \begin{table} \caption{ Prediction for the rare radiative decay form factor $F(0) $ in comparison with various theoretical predictions.} \vspace {0.2 true in} \begin{tabular}{l|l|l|l|l} \hline \hline \multicolumn{1}{l|}{Decay process}& \multicolumn{1}{|l|}{Present work} & \multicolumn{1}{|l|}{Ref. \citen{ref8}}& \multicolumn{1}{|l|}{Ref. \citen{ref31}} & \multicolumn{1}{|l}{Ref. \citen{ref32}}\\ \hline & & & & \\ $B^{\pm} \rightarrow K^{* \pm} \gamma $ & 0.328 &- &- &- \\ & & & & \\ $B^{0} \rightarrow K^{*0} \gamma$ & 0.329 & $0.32 \pm 0.03$& $0.32 \pm 0.05 $ &$0.308 \pm 0.013 \pm 0.036 \pm 0.006 $\\ & & & & \\ $B^- \rightarrow \rho^{-} \gamma$ & 0.295 & $ 0.26 \pm 0.03 $ & $ 0.24 \pm 0.04 $ & $0.27 \pm 0.011 \pm 0.032$\\ & & & & \\ $B_s \to K^{* 0} \gamma $ & 0.250 & $0.23\pm 0.02 $ & $ 0. 20 \pm 0.04 $& -\\ &&& & \\ \hline \end{tabular} \end{table} \begin{table} \caption{ Prediction for the branching ratios of the exclusive rare decays of $B$ mesons along with their experimental values.} \vspace {0.2 true in} \begin{tabular}{l|l|l} \hline \hline \multicolumn{1}{l|}{Decay process}& \multicolumn{1}{|l|}{Br ratio} & \multicolumn{1}{|l}{Br ratio}\\ \multicolumn{1}{l|}{} & \multicolumn{1}{|l|}{This work}& \multicolumn{1}{|l}{Expt. \citen{ref19}}\\ \hline & & \\ $B^0 \rightarrow K^{*0} \gamma $ & 3.96 $\times 10^{-5}$ & $ (4.0 \pm 1.9) \times 10^{-5}$ \\ & & \\ $B^{\pm} \rightarrow K^{*\pm} \gamma$ & 4.168 $\times 10^{-5}$ & $(5.7 \pm 3.3)\times 10^{-5}$ \\ & & \\ $B^- \rightarrow \rho^{-} \gamma$ & $1.68 \times 10^{-6}$ &- \\ & & \\ $B_s \to K^{* 0} \gamma $ & $1.158 \times 10^{-6} $ & -\\ & & \\ \hline \end{tabular} \end{table} \newpage \acknowledgements R. M. would like to thank CSIR, Govt. of India, for a fellowship. A. K. G. and M. P. K. acknowledge financial support from DST, Govt. of India.
\section{Introduction and motivation} \label{sec:intro} `If Supersymmetry (SUSY) exists, it will be discovered at the next generation of hadronic machines', has been a recurring motto so far. Indeed sooner (at the Tevatron, $\sqrt s=2$ TeV) or later (at the Large Hadron Collider (LHC), $\sqrt s=14$ TeV), depending on the mass scale of the Higgs bosons and of the Superpartners of ordinary matter, several Supersymmetric `signatures' should clearly be viable\footnote{For some reviews, see, e.g., \cite{Tevareview} and \cite{LHCreview}.}. Typical SUSY events at hadron colliders will involve either the production and decay of heavy spartons, squarks and gluinos, whose foreseen mass range is expected to be around the TeV scale \cite{signals}, or of Higgs bosons \cite{Higgs}, primarily of the lightest one, for which the SUSY theory imposes a stringent mass bound of the order of the electroweak (EW) scale. However, even assuming that such a discovery will take place, there might well be little to learn about the fundamental dynamics of SUSY from such new events. In fact, although, e.g., the LHC is able to produce gluinos and squarks with masses up to 2 TeV or so and their detection has been shown to be feasible with rather little effort \cite{HPSSY}, it is much more difficult to determine exactly the SUSY masses involved, because in most models (i.e., those assuming $R$-parity conservation) there are at least two missing SUSY particles in each event. Clearly, failing the knowledge of the SUSY mass spectrum, other typical SUSY quantities, such as couplings, decay rates, etc., cannot be assessed either. Needless to say, their measurement would be of paramount importance in order to constrain the free parameters entering the SUSY Lagrangian. However, by resorting to specific kinematic distributions \cite{HPSSY}, it is at least possible to make precision measurements of some `combinations' of SUSY masses, but only in a few fortunate cases these can lead to strong constraints on the theory and its parameters. Besides, in minimal SUSY theories, the Higgs sector typically (i.e., at tree level) depends on only two such parameters, the ratio of the vacuum expectation values (VEVs) of the Higgs fields and one of the masses of the five physical states corresponding to the latter, as all others SUSY inputs enter through higher perturbative orders. Therefore, even the detection of a SUSY Higgs signal would carry very poor information in terms of the underlying SUSY model. As a matter of fact, a second question about SUSY has to legitimately be risen. Namely, `Which Supersymmetric model will one discover ?' Thus, the key task for the Tevatron and the LHC is {not} only to find SUSY, but also to assess which model is behind it and the value of its parameters. For example, in the context of Supergravity (SUGRA) inspired models \cite{SUGRA}, with the minimal particle content of the MSSM (henceforth denoted as M-SUGRA, that we take to be the reference framework of our analysis) \cite{MSSM,MSUGRA}, the dynamics of the theory can be specified by only five entries: (i) a universal scalar mass $M_0$; (ii) similarly, that for the gauginos $M_{1/2}$; (iii) the universal trilinear breaking terms $A_0$ (all defined at the Grand Unification scale $M_{\mathrm{GUT}}$ \cite{graham}). After the radiative EW symmetry breaking has taken place, two further parameters are needed to describe the low energy dynamics: (iv) the mentioned ratio of the VEVs of the two Higgs fields, denoted by $\tan\beta\equiv v/v'$ and defined at the EW scale; and (v) a discrete parameter, ${\mbox{sign}}(\mu)= \pm1$, being $\mu$ the Higgsino mass term. Assuming universal soft breaking terms at the GUT scale, one is then able to calculate the masses of SUSY (s)particles, their couplings, decay rates, etc., at the EW scale, through the evolution of the renormalisation group equations (RGEs), the latter involving $M_0$, $M_{1/2}$, $A_0$, $\tan\beta$ and ${\mbox{sign}}(\mu)$ as inputs. Ultimately, a comparison of such predictions with the corresponding experimental measurements, as reconstructed from the actual data via dedicated Monte Carlo (MC) simulations \cite{isajet}--\cite{herwig}, should allow one to impose indirect constraints on the above parameters. Indeed, an additional procedure to follow in order to determine the latter could well be to search for the evidence of some more exotic signals of SUSY, in which, however, the dependence on such parameters is somewhat more manifest. Elementary processes of the type \begin{equation}\label{proc} g + g \longrightarrow {\tilde{q}}_{\chi} + {\tilde{q}}^{'*}_{\chi'} + \Phi, \end{equation} where $q^{(')}=t,b$, $\chi^{(')}=1,2$ and $\Phi=H,h,A,H^\pm$, in all possible combinations, as appropriate in the MSSM, serve the double purposes of: \begin{enumerate} \item furnishing production mechanisms of Higgs bosons of the MSSM, both neutral and charged, in addition to the Standard Model (SM)-like channels \cite{Higgs}; \item yielding production rates, for particular combinations of $q^{(')}$, $\chi^{(')}$ and $\Phi$, strongly dependent on some of the fundamental SUSY parameters of the M-SUGRA model. \end{enumerate} The importance of the first point should be understood in the following terms. On the one hand, the detection of all neutral Higgs particles $H,h$ and $A$ of the theory is not certain, neither at the Tevatron \cite{Tevareview} nor even at the LHC \cite{LHCreview}. In addition, the discovery potential of heavy charged Higgs bosons $H^\pm$ at both the above colliders has been proved to be extremely limited \cite{charged}. Under these circumstances, the possible existence of novel and detectable Higgs production channels represents a phenomenologically important result per se. (Notice that, for certain choices of $\mathrm{sign}(\mu)$, $A_0$ and $\tan\beta$, the squark-squark-Higgs coupling can become the largest EW coupling of the SUSY theory, even exceeding the standard Yukawa ones.) On the other hand, the fact that in processes (\ref{proc}) the Higgs bosons are produced in association with squarks via a Yukawa bremsstrahlung or in `non-dominant' squarks decays\footnote{So that the corresponding partial widths are significantly different from the total widths, thus retaining the dynamics of the squark-squark-Higgs production vertices also at decay level.}, implies that the Higgs mechanism can be probed in the sparticle sector too. In fact, other known production and decay mechanisms used to detect MSSM Higgs bosons mainly involve Higgs couplings to ordinary matter. The only exceptions are the squark loop-contributions to neutral Higgs boson production via gluon-gluon fusion and to Higgs boson decays through pairs of photons/gluons \cite{xggh}, which are however swamped in both modes by the dominant terms involving ordinary heavy particles. As for the second point that we put forward, we should remind the reader of the actual form of the mentioned squark-squark-Higgs vertices in the MSSM (which can be found, e.g., in \cite{HHG}). In many of these, namely when $\chi\ne\chi'$, {both} the low-energy SUSY parameters $\mu$ and $\tan\beta$ enter {explicitly} in the Feynman rules, other than {implicitly} in the scalar masses. Furthermore, those vertices also contain $A_{{{q}}^{(')}}$, the trilinear couplings at the EW scale, which depend critically on their common value at the GUT scale, $A_0$. (In the case $\Phi=A$ such a dependence is also not affected by the mixing between the chiral, $\chi^{(')}=L,R$, and physical, $\chi^{(')}=1,2$, squark states, so that no additional SUSY mixing parameters enter the phenomenology of pseudoscalar Higgs production, this rendering the latter an ideal laboratory to study M-SUGRA effects \cite{short}). Therefore, one should expect a significant dependence of the production rates of the scattering processes (\ref{proc}) on $\tan\beta$, $A_0$ and $\mu$ (particularly, its sign), this possibly yielding a new profitable mean to constrain the underlying SUSY model. Even more so in the case $\tan\beta$ has previously been determined, for example, through a discovery in the MSSM Higgs sector. Concerning previous literature on the subject, we should mention that reactions of the type (\ref{proc}) were were first considered in Ref.~\cite{djouadi} for the case $g g \rightarrow \tilde{t}_1 \tilde{t}_1^* h$ in the so-called `decoupling' limit. Adopting the M-SUGRA scenario, associated production of both neutral and charged Higgs bosons production with squark pairs -- with a special emphasis on CP-odd Higgs boson production -- was first consider by the authors in \cite{short}. Furthermore, in Ref.~\cite{djouadi2} (and also \cite{fawzi}) light Higgs boson production in association with light top squarks was reanalysed in the M-SUGRA scenario at both Tevatron and LHC\footnote{The same final state but produced in $e^+e^-$ annihilations at the TeV scale, e.g., at the Next Linear Collider, has been considered in Refs. \cite{djouadi2,ee}.}. A general consensus on the possible detectability at the LHC of $gg\to {\tilde{t}}_1{\tilde{t}}_1^* h$ events emerged from Refs.~\cite{djouadi,djouadi2,fawzi}, in the case of light top squarks and large trilinear coupling. We generalise here those studies, as we consider the production of all possible Higgs states, $\Phi=H,h,A$ and $H^\pm$, for a broader spectrum of their masses, in conjunction with both squark flavours that can have sizable couplings to Higgs particles, ${\tilde{q}}={\tilde{t}}$ and ${\tilde{b}}$, the latter taken as not degenerate in mass, also allowing for Higgs production in decay channels, when $\chi\ne\chi'$. On the other hand, to simplify the simulation, we restrict ourselves to the case of gluon-gluon induced processes only, thus neglecting the case of quark-antiquark scatterings, which was instead considered in Refs.~\cite{djouadi,djouadi2}. This is however not restrictive. In fact, we have verified that at the LHC the $gg$ contribution is around two order of magnitudes larger than the $Q\bar Q$ one, in line with the findings of Refs.~\cite{djouadi,djouadi2}, well below the level of uncertainties arising in our computation from other sources (such as structure functions, QCD $K$-factors, etc.)\footnote{As for $\gamma,Z$ and $W^\pm$ $s$-channels, these are typically smaller by a factor of the order ${\cal O}({\alpha_{\mathrm{em}}}/{\alpha_{\mathrm{s}}})^2$; whereas ${\cal O}({\alpha_{\mathrm{s}}})^2$ gluino interactions in $t,u$-channels are suppressed by the small (EW induced) mixing between light quarks and sbottom and stop squarks, as already remarked in \cite{djouadi}.}. As for the Tevatron, we can anticipate that the production cross sections of processes (\ref{proc}) are generally very small (see also Ref.~\cite{djouadi2}), indeed below the level of detection over most of the M-SUGRA parameter space, so that we neglect further consideration of this machine here. It is the purpose in this paper to assess the possible relevance of points 1. and 2. in phenomenological studies of SUSY to be carried out at the LHC. In particular, the plan of the paper is as follows. In the next Section we describe how we have proceeded in our calculations. In Sect.~\ref{sec:parameters} we illustrate the theoretical model we have resorted to in our analysis. Sect.~\ref{sec:results} presents some numerical results. A brief summary and our conclusions are given in Sect.~\ref{sec:conclusions}. Finally, we collect in Appendix the relevant analytical formulae that we have used. \section{Calculation} \label{sec:calc} The leading-order (LO) Feynman diagrams associated to processes of the type (\ref{proc}) in the unitary gauge are depicted in Fig.~\ref{fig:graphs}. The reader can find an analytical expression of the corresponding matrix elements (MEs) in the Appendix. As a test of the correctness of our amplitudes, we have verified that they are gauge invariant by checking various Ward identities of the theory, both analytically and numerically. The amplitudes squared have then been integrated over a three-body phase space, using VEGAS \cite{VEGAS}, and convoluted with gluon Parton Distribution Functions (PDFs), as provided by CTEQ(4L) \cite{CTEQ4}. The latter constitutes our default set, taken at LO in order to be consistent with our approximation in calculating the scattering MEs. However, in order to estimate the systematic error due the gluon behaviour inside the proton, we also have resorted to other LO packages, such as MRS-LO(09A,10A,01A,07A) \cite{MRS98LO}. Typical differences among PDFs were found to be less than 15--20\%. The centre-of-mass (CM) energy at the partonic level, $Q=\sqrt{\hat{s}}$, was the scale used to evaluate both the structure functions and the strong coupling constant (see next Section for the treatment of the latter). Depending on the relative value of the final state masses in (\ref{proc}), $m_{{\tilde{q}}_{\chi}}$, $m_{{\tilde{q}}^{'*}_{\chi'}}$ and $m_{\Phi}$, the production of Higgs particles can be regarded as taking place either via a (anti)squark decay (if $m_{{\tilde{q}}_{\chi}} > m_{{\tilde{q}}^{'*}_{\chi'}} + m_{\Phi}$ or $m_{{\tilde{q}}^{'*}_{\chi'}} > m_{{\tilde{q}}_{\chi}} + m_{\Phi}$) or via a Higgs-strahlung (if $m_{{\tilde{q}}_{\chi}} < m_{{\tilde{q}}^{'*}_{\chi'}} + m_{\Phi}$ and $m_{{\tilde{q}}^{'*}_{\chi'}} < m_{{\tilde{q}}_{\chi}} + m_{\Phi}$). In the first case, to prevent our MEs from becoming singular, we need to insert a finite width in the resonant (anti)squark propagators, which we have done by adopting the Breit-Wigner expression given in the Appendix and the appropriate numerical values for the widths, calculated as described in Sect.~\ref{sec:parameters}. Also in the second case, though no poles exist in the amplitudes, a finite width value has been retained in the propagators. Notice that we have treated the two processes on the same footing, without making any attempt to separate them, as for the time being we are only interested in the total production rates of the 2 $\to$ 3 processes (\ref{proc}), rather than in their subsequent decay distributions. \section{The theoretical model and its parameters} \label{sec:parameters} In this paper we are going to display our results for squark-squark-Higgs production via processes (\ref{proc}) by assuming possibly the simplest scenario in the choice of the soft SUSY breaking parameters at the GUT scale. That is, the so-called minimal Supergravity scenario or M-SUGRA inspired model, as already intimated in the Introduction. In this scenario, the whole dynamics of the MSSM (which contains over hundred parameters in the case of conserved $R$-parity) at the GUT scale is reduced to the three basic inputs already introduced: $M_0$, $M_{1/2}$ and $A_0$. The large top Yukawa coupling then triggers the radiative EW breaking through the running of the soft Higgs breaking masses, from the GUT scale down to EW regime. {}From the minimisation conditions of the potential one can define the soft Higgs mixing parameter, $B$, and the absolute value of the Higgs mixing parameter of the SUSY potential, $\mu$. The model leaves the sign($\mu$) and the value of $\tan\beta$ as further undetermined parameters at the EW scale. All five M-SUGRA parameters enter into the relevant Feynman rules for the squark-squark-Higgs vertices, either explicitly or implicitly (through the RGEs). These can be written in the physical squark basis $\tilde{q}_{1,2}$ as \begin{eqnarray}\label{mixing} \lambda_{\Phi\tilde{q}_1\tilde{q}'_1} &=& c_q c_{q'} \lambda_{\Phi \tilde{q}_L\tilde{q}'_L} + s_q s_{q'} \lambda_{\Phi\tilde{q}_R\tilde{q}'_R}+ c_q s_{q'} \lambda_{\Phi\tilde{q}_L\tilde{q}'_R} + s_q c_{q'} \lambda_{\Phi\tilde{q}_R\tilde{q}'_L} ,\nonumber \\[2mm] \lambda_{\Phi\tilde{q}_2\tilde{q}'_2} &=& s_q s_{q'} \lambda_{\Phi \tilde{q}_L\tilde{q}'_L} + c_q c_{q'} \lambda_{\Phi\tilde{q}_R\tilde{q}'_R} -s_q c_{q'} \lambda_{\Phi\tilde{q}_L\tilde{q}'_R} -c_q s_{q'} \lambda_{\Phi\tilde{q}_R\tilde{q}'_L} ,\nonumber \\[2mm] \lambda_{\Phi\tilde{q}_1\tilde{q}'_2} &=& -c_q s_{q'} \lambda_{\Phi \tilde{q}_L\tilde{q}'_L} + s_q c_{q'} \lambda_{\Phi\tilde{q}_R\tilde{q}'_R} +c_q c_{q'} \lambda_{\Phi\tilde{q}_L\tilde{q}'_R} -s_q s_{q'} \lambda_{\Phi\tilde{q}_R\tilde{q}'_L} ,\nonumber \\[2mm] \lambda_{\Phi\tilde{q}_2\tilde{q}'_1} &=& -s_q c_{q'} \lambda_{\Phi \tilde{q}_L\tilde{q}'_L} + c_q s_{q'} \lambda_{\Phi\tilde{q}_R\tilde{q}'_R} -s_q s_{q'} \lambda_{\Phi\tilde{q}_L\tilde{q}'_R} +c_q c_{q'} \lambda_{\Phi\tilde{q}_R\tilde{q}'_L} , \label{fr} \end{eqnarray} where $\tilde{q}_{L,R}$ or $\tilde{q}'_{L,R}$ can in principle be any flavour of chiral squarks. However, here we only focus our attention to the case of the the third generation of down and up squarks only, namely, sbottom and stop scalars, whose physical mass eigenstates are denoted by $\tilde{b}_{1,2}$ and $\tilde{t}_{1,2}$, respectively, the subscript 1(2) referring to the lightest(heaviest) of them. As usual, the Higgs fields are denoted by the generic symbol $\Phi$, where $\Phi=H,h,A,H^\pm$. All the $\lambda_{\Phi\tilde{q}_\chi\tilde{q}'_{\chi'}}$'s appearing in eq.~(\ref{mixing}) are function of $\mu, \tan\beta$ and $A_{t,b}$ and can be read directly from the Appendix of Ref.~\cite{HHG}. (We ignore the case of complex $\mu$ and $A_{q}$ ($q=t,b$) parameters by assuming that their phases are very small, the preferred case following the measurements of the Electric Dipole Moments \cite{nir}.) Also the left-right squark mixing angles $s_q\equiv\sin\theta_q$ and $c_q\equiv\cos\theta_q$ (here, $q=t,b$) depend on the M-SUGRA parameters, since they read as \begin{eqnarray}\label{angle} \tan(2\theta_t) &=& \frac{2 m_t (A_t - \mu \cot\beta)}{M_{\tilde{Q}_3}^2- M_{\tilde{U}_3^c}^2+(\frac{1}{2}-\frac{4 s_W^2}{3})M_Z^2 \cos2\beta}, \\[2mm] \nonumber \tan(2\theta_b) &=& \frac{2 m_b (A_b - \mu \tan\beta)}{M_{\tilde{Q}_3}^2- M_{\tilde{D}_3^c}^2+(-\frac{1}{2}+\frac{2 s_W^2}{3})M_Z^2 \cos2\beta}, \end{eqnarray} with $M_Z$ the $Z$-boson mass and $s_W\equiv\sin^2\theta_W$ the sine (squared) of the Weinberg angle, $m_t$ and $m_b$ the top and bottom quark masses, where $A_t$ and $A_b$ are the trilinear couplings defined at the EW scale, while $M_{\tilde{Q}_3}$, $M_{\tilde{U}_3^c}$ and $M_{\tilde{D}_3^c}$ are the running soft SUSY breaking squark masses of the third generation, for which we assume the values obtained from their evolution starting from a universal mass at the GUT scale equal to $M_0$. Regarding the numerical values of the M-SUGRA parameters adopted in this paper, we have proceeded as follows. For a start, we have set $M_0=M_{1/2}=150$ GeV. Such rather low values for the universal masses come as natural first choice, if one is interested in detecting processes of the type (\ref{proc}). For two simple reasons. On the one hand, these two quantities determine the actual $m_{\tilde q_\chi}$, $m_{\tilde q'_{\chi'}}$ and $m_\Phi$ values entering processes (\ref{proc}), through their intervention in the RGEs, in such a way that small values of $M_0$ and $M_{1/2}$ at the GUT scale convert into a rather light squark and Higgs mass spectrum at the EW scale. On the other hand, being $2\to3$ body processes, a strong suppression from the phase space would arise in squark-squark-Higgs production if the masses in the final state were too large\footnote{The additional depletion coming from the gluon PDFs, which would be probed at much higher values of $Q^2$ (of the order of the rest masses or more), where they are naturally smaller, would in part be compensated by the rise of the quark-antiquark initiated subprocesses: i.e., $Q+\bar Q\rightarrow {\tilde{q}}_{\chi} + {\tilde{q}}^{'*}_{\chi'} + \Phi$, where, again, one has that $q^{(')}=t,b$, $\chi^{(')}=1,2$ and $\Phi=H,h,A,H^\pm$.}. For the above choice, the M-SUGRA model predicts squark and heavy Higgs masses in the region of 80--450 GeV (as we shall see in more detail below), so that the latter can in principle materialise at LHC energies (further recall that the light Higgs mass is bound to be below 130 GeV). Then, we have varied the trilinear soft Supersymmetry breaking parameter $A_0$ in a region where it changes its sign, e.g., $(-300,+300)$ GeV, while we spanned the $\tan\beta$ value between 2 and 40. As for $\mu$, whereas in our model its magnitude is constrained, its sign is not. Thus, in all generality, we have explored both the possibilities $\mathrm{sign}(\mu)=\pm1$. As a further step of our analysis, we have then come back to $M_0$ and $M_{1/2}$ and change them, while maintaining $\tan\beta$, $A_0$ and sign$(\mu)$ fixed at some specific values. We have done so only for those processes that we had already identified to have not only a large cross section, but also a strong dependence on one or more of these three M-SUGRA parameters. Given the strong phase space suppression induced by the consequent increase of $m_{\tilde{q}_\chi}$, $m_{\tilde{q}'_{\chi'}}$ and $m_\Phi$ in the final states, we will cautiously maintain the universal scalar and gaugino masses below 250--300 GeV (at least at first). However, the reader should not assume that this is a necessary condition to the experimental detection of processes of the type (\ref{proc}). In fact, this need not be true, as we shall show that even for $M_0$ values as large as 500 GeV one can find sizable squark-squark-Higgs production cross sections for $M_{1/2}$ up to 200--250 GeV, as long as $A_0$ is strongly negative, $\tan\beta\buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} 30$ and sign$(\mu)<0$. Such unexpected behaviours are strongly driven by the intervention in the production rates of $\tan\beta$, $A_0$ and $\mathrm{sign}(\mu)$, through the trilinear scalar couplings $\lambda_{\Phi\tilde{q}_\chi\tilde{q}'_{\chi'}}$, more than by the actual values of $m_{\tilde{q}_\chi}$, $m_{\tilde{q}'_{\chi'}}$ and $m_\Phi$. This is evident by a mere look at the standard Feynman rules, as can be found, e.g., in Ref.~\cite{HHG}. We will make our concern in this paper that of guiding the reader through such delicate interplay between masses and couplings, by explicitely writing down the expression of the relevant vertices in those parameter space domains where such complicated phenomenology manifests itself. {Starting from the five M-SUGRA parameters $M_0$, $M_{1/2}$, $A_0$, $\tan\beta$ and ${\mbox{sign}}(\mu)$, we have generated the spectrum of masses, widths, couplings and mixings relative to squarks and Higgs particles entering reactions (\ref{proc}) by running the {\tt ISASUGRA/ISASUSY} programs contained in the latest release of the package {\tt ISAJET} \cite{isajet}, version 7.40. The default value of the top mass we used was 175 GeV. Note that also typical EW parameters, such as $\alpha_{\mathrm{em}}$ and $\sin^2\theta_W$, were taken from this program, as they enter the RGEs of the SUSY theory. Concerning the value of the strong coupling constant, $\alpha_{\mathrm{s}}$, entering the production processes (\ref{proc}), we have proceeded as follows. By using as inputs the extracted value of $\alpha_{\mathrm{s}}$ at the $M_Z$ scale, we evolve it up to any scale $Q$ by making use of the two-loop renormalisation group equations and by taking into account all the low-energy threshold effects from the various SUSY masses by means of the theta function approximation, as discussed in Ref. \cite{sakis}. Finally, notice that in scanning over the M-SUGRA parameter space, one should make sure that the values generated for the Higgs boson and sparticle masses are in accordance with current experimental bounds. Signs of the sort ``$\times$'' or shaded areas appearing in our figures in forthcoming Sect.~\ref{sec:results} will correspond to already prohibited areas for the parameter space of our SUSY model. We nonetheless leave them for illustrative purposes, in order to visualise the typical impact of present and future experimental bounds on the phenomenology of our reactions. For example, the M-SUGRA points individuated by the combinations $M_0=M_{1/2}=130-150$ GeV, $\tan\beta=2$, $A_0=0$ GeV and $\mathrm{sign}(\mu) = -$, used in some of the tables and figures in the next Section, contradict the limits on the lightest Higgs boson mass from direct searches \cite{lep,ALEPH}, as they yield $m_h=72-80$ GeV. We will discuss the experimental bounds in more detail in the next Section. As for theoretical constrains, these arise from two sources: namely, the absence of charge and colour breaking minima and that of large contributions to the EW observables that are measured with high precision at LEP. The former is avoided when the following inequalities hold tree level~\cite{lahanas}: \begin{eqnarray} A_t^2 \ &<& \ 3 \left (m_{\tilde{Q}_3}^2+m_{\tilde{U}^c_3}^2+\mu^2 +m_{H_2}^2 \right ) , \nonumber \\ A_b^2 \ &<& \ 3 \left (m_{\tilde{Q}_3}^2+m_{\tilde{D}^c_3}^2+\mu^2 +m_{H_1}^2 \right ) , \nonumber \\ A_{\tau}^2 \ &<& \ 3 \left (m_{\tilde{L}_3}^2+m_{\tilde{E}^c_3}^2 +\mu^2 +m_{H_1}^2 \right ) , \label{ccb} \end{eqnarray} where all masses appearing in eq.~(\ref{ccb}) are the soft Supersymmetry breaking masses except the Higgsino mixing parameter $\mu$. When $A_0$ is below 1 TeV, as is the case in our analysis, the above constraints are always satisfied even for very light squark masses. As for the contributions to the EW observables, we have found the region $150~{\mathrm{GeV}}\buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} M_0,M_{1/2} \buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} 500$ GeV, $2\buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} \tan\beta \buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} 40$ and $|A_0| \buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} 900$ GeV covered by our analysis in agreement with the most recent measurements of the `effective' weak mixing angle, $\sin^2\theta_{\mathrm{eff}}=0.2321\pm 0.010$, and of the $W^\pm$ mass, $M_W=80.388\pm 0.063$ GeV \cite{erler}. In particular, notice that the M-SUGRA prediction for $\sin^2\theta_{\mathrm{eff}}$ decreases for a lighter mass spectrum while it becomes constant in the heavy mass region \cite{sakis1}. \section{Results} \label{sec:results} We begin this Section by analysing all reactions (\ref{proc}) in the low mass regime, i.e., that induced by values of $M_0$ and $M_{1/2}$ below 250--300 GeV. This is done in Subsect.~\ref{lightspectrum}. In this scenario, we will first present and discuss, for future reference, the values of squark and Higgs masses resulting from the RGE evolution: in \ref{squarks} and \ref{higgses}, respectively. Then, we will move on to considering the production of, in turn: neutral CP-even (in \ref{even}), CP-odd (in \ref{odd}) and charged (in \ref{charged}) Higgs bosons. Possible decay signatures of the latter will be analysed in \ref{signatures}. Finally, Subsect.~\ref{heavyspectrum} will pin-point those unusual cases discussed above, in which the suppression from very heavy scalar masses in the final states of reactions (\ref{proc}) can be overcome by strong vertex effects, yielding in the end sizable cross sections. \subsection{Light mass spectrum} \label{lightspectrum} Once fixed $M_0=M_{1/2}=150$ GeV, one obtains the (sbottom and stop) squark and Higgs masses reported in Figs.~\ref{fig:squarks} and \ref{fig:higgs}, respectively, depending on 2 $\buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim}\tan\beta\buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim}$ 35 and for $A_0=-300,0,+300$~GeV. The two possible options for the sign of $\mu$ are also contemplated. Far from willing to discuss exhaustively the dependence of $m_{\tilde q_\chi}$, $m_{\tilde q'_{\chi'}}$ and $m_\Phi$ upon the five M-SUGRA parameters, we limit ourselves here to spotting in Figs.~\ref{fig:squarks}--\ref{fig:higgs} some interesting trends, that will affect the overall behaviour of the squark-squark-Higgs cross sections that we will be treating below. For a more complete overview, see, e.g., Refs. \cite{pokorski}. \subsubsection{Squark masses} \label{squarks} The four squark flavours of the first and second generation ($\tilde{q}= \tilde{u} , \tilde{d} , \tilde{c} , \tilde{s}$) with left- and right-handed components are all nearly degenerate in mass and the latter is given approximately by the following formula: \begin{equation} m_{\tilde{q}} \simeq \sqrt{M_0^2 + 6 M_{1/2}^2 }, \qquad\qquad\qquad (\tilde{q}= \tilde{u} , \tilde{d} , \tilde{c} , \tilde{s}). \end{equation} For our choice of $M_0$ and $M_{1/2}$, one gets $m_{\tilde q}\simeq 400$ GeV. The light squark flavours are not our concern in processes (\ref{proc}), though they may well enter some of the decay chains of the other (pseudo)scalar particles produced. The two squark flavours of the third generation must be treated differently because the off-diagonal entries of their mass matrices can be large, owing to the strength of the Yukawa couplings of the corresponding quarks (in this respect, notice that the bottom one becomes comparable to that of the top in the large $\tan\beta$ region \cite{HHG}). It thus follows that the mass eigenstates $\tilde{t}_1$, $\tilde{t}_2$, $\tilde{b}_1$ and $\tilde{b}_2$ are all different and generally smaller than $m_{\tilde q}$. Among these, $\tilde{t}_1$ is most often significantly lighter than all the other stop and sbottom states. At large $\tan\beta$, the same happens to the $\tilde{b}_1$ mass eigenstate, as large values of $\tan\beta$ correspond to smaller sbottom masses, so that one eventually gets that $m_{\tilde{t}_1} \simeq m_{\tilde{b}_1}$. Variation of the trilinear couplings can also cause significant differences between the light stop and sbottom masses. Finally, the $\mathrm{sign}$ of the Higgsino mass term plays an important r{\^o}le when $\tan\beta$ is either small or large, for the cases $m_{\tilde{t}_{1,2}}$ and $m_{\tilde{b}_{1,2}}$, respectively. Experimental limits on the squark masses come from searches at Tevatron and LEP2. The most stringent bound on the ${{\tilde t}_1}$ mass comes from the hadron collider \cite{Tevatron}: in absence of mixing, values of $m_{{\tilde t}_1}<120$ GeV are excluded for\footnote{Recall that in M-SUGRA the Lightest Supersymmetric Particle (LSP) is the lightest neutralino, ${\tilde\chi}_1^0$.} $m_{\mathrm{LSP}}\equiv m_{\tilde{\chi}_1^0} <$ 38 GeV. D\O ~exclude values of $m_{\tilde{b}_1}$ below 85 GeV for $m_{\tilde{\chi}_1^0} <$ 47 GeV \cite{d0} and ALEPH do over the region $m_{\tilde{b}_1}<83$ GeV for any value of LSP mass \cite{aleph1}. In addition, CDF exclude masses for the lightest top squark up to 120 GeV when the LSP is $m_{\tilde{\chi}_1^0} <$ 50 GeV \cite{Tevatron}. Finally, D\O, using data corresponding to an integrated luminosity of 79 pb$^{-1}$, contradicts all models with $m_{\tilde{q}\ne\tilde{t}_1,\tilde{b}_1 }<$ 250 GeV for $\tan\beta\Ord2$, $A_0=0$ GeV and $\mu< 0$ \cite{d01} (in scenarios with equal squark and gluino masses the limit goes up to $m_{\tilde{q}\ne\tilde{t}_1,\tilde{b}_1}<260$ GeV). \subsubsection{Higgs boson masses} \label{higgses} The mass of the lightest Higgs boson is here constrained to be less than 115 GeV and it appears to exhibit constant values over the region $8\buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim}\tan\beta \buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} 35$, for a given combination of $A_0$ and $\mathrm{sign}(\mu)$. To change either the trilinear coupling or the sign of the Higgsino mass has the net effect of scaling $m_h$, lower with increasing $A_0$ and higher for positive $\mu$. As for the masses of the other Higgs bosons, they are nearly degenerate. They vary between 180 GeV (when $\tan\beta$ is large) and 400 GeV (when $\tan\beta$ is small). The dependence on $\mathrm{sign}(\mu)$ is generally negligible, whereas the one on $A_0$ is very strong in the intermediate $\tan\beta$ regime. One thing worth noticing here is the existence of a point where all three Higgs masses $m_H,m_A$ and $ m_{H^\pm}$ converge, regardless of the values of $A_0$ and $\mathrm{sign}(\mu)$. This occurs for $\tan\beta \approx 30$, where \begin{equation} m_H=m_A=m_{H^\pm} \approx 200~{\rm GeV}. \end{equation} As for experimental bounds, LEP experiments have combined their results from data taken at CM energies from 91 to 183 GeV to place lower bounds on the masses of the light ($m_h$) and pseudoscalar ($m_A$) Higgs bosons, of 78.8 and 79.1 GeV, respectively \cite{hocker}. In addition, they exclude the range $0.8<\tan\beta <2.1$ for minimal stop mixing and $m_t=175$ GeV. Also, D\O\ \cite{d02} have recently removed at 95\% confidence level the intervals $\tan\beta <0.97$ and $\tan\beta > 40.9$ for $M_{H^\pm}=60$ GeV and $\sigma(t\bar{t})=5.5$ pb (again, with $m_t=175$ GeV). However, the limits become less stringent with increasing $M_{H^\pm}$: e.g., for $M_{H^\pm}>124$ GeV (as is the case here) the available angular range is $0.3<\tan\beta<150$. \subsubsection{CP-even Higgs boson production} \label{even} We present in Figs.~\ref{fig:lighthiggs}--\ref{fig:heavyhiggs} the $\tan\beta$ dependence of the production cross sections of the two neutral scalar Higgs bosons, $h$ and $H$, respectively, in association with any possible combination of squarks of the third generation. Again, we parametrise the dependence upon $A_0$ by adopting for the latter the discrete values of $-300$, 0 and +300 GeV and we choose $\mathrm{sign}(\mu)=\pm1$. Assuming an integrated luminosity of 100 inverse femtobarns over a twelve month period of running at the LHC (i.e., $\int {\cal L} dt = 100 \; {\mathrm{fb}}^{-1}$), one can realistically hope for the detection of squark-squark-Higgs processes only if the production cross section is above 1 fb or so. In fact, we shall see in Sect.~\ref{signatures} how typical decay fractions of clean signatures range at the level of 10\% or below (see Subsect.~\ref{signatures} later on). Under this assumption, one immediately sees that there are several production channels of CP-even Higgs particles which could be observed, over a large part of the M-SUGRA parameter space considered here. Primarily, those involving the lightest stop squark, ${\tilde t}_1$, particularly if also the lightest Higgs state is involved, but not only. For the case of $h$ production, there exists an approximate hierarchy of cross sections which can possibly be detected: \begin{equation}\label{Xsections_h} \sigma( \tilde{t}_1 \tilde{t}_1^* h )\buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} \sigma( \tilde{t}_1 \tilde{t}_2^* h )\buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} \sigma( \tilde{b}_1 \tilde{b}_1^* h ). \end{equation} The cases $\tilde{b}_2 \tilde{b}_2^* h$, $\tilde{t}_2 \tilde{t}_2^* h$ and $\tilde{b}_1 \tilde{b}_2^* h$ never have significantly large rates. In the case of $\tilde{t}_1 \tilde{t}_1^* h $ states, those with largest production rates, one obtains \begin{equation} \sigma (g g \rightarrow h \tilde{t}_1 \tilde{t}_1^* ) \buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} 20 \; {\rm fb}, \end{equation} for every combination of $\tan\beta$, $A_0$ and $\mathrm{sign}(\mu)$, in the case of $M_0=M_{1/2}=150$ GeV, thus a sort a lower limit over a representative portion of the low mass regime of the M-SUGRA scenario. Moreover, the largest production rate for this final state (compatible with the current experimental constraints) is obtained in the small $\tan\beta \sim 2 $ region and for the combinations $A_0\buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} -300$ and $\mu =+$, for which \begin{equation} \sigma ( \tilde{t}_1 \tilde{t}_1^* h) \buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} 200 \; {\rm fb}, \end{equation} corresponding to more than 20000 events per year running of the LHC. The dominance of the production channel involving both the lightest squarks and Higgs boson, above all other mechanisms (\ref{proc}), was foreseeable. The reason is rather simple, in fact, twofold. On the one hand, the sum of the rest masses in the final states yields the smallest possible values, thus enhancing the volume of the three-body phase space, relatively to any other squark-squark-Higgs combination. On the other hand, the cross section is also significantly enhanced when the trilinear coupling $A_0$ assumes negative values. From the analysis of the RGEs we find that $A_t \sim -400 (-250)$ GeV when $A_0=-300 (300)$ GeV. Now, the coupling of the light Higgs boson to top squarks is driven by $A_t$ when the latter takes on large values. More specifically, the corresponding vertex reads as (recall that we are in the kinematic limit $m_{X}\approx m_H\approx m_A\approx m_{H^\pm}\gg M_Z$: see Fig.~\ref{fig:higgs}) \begin{eqnarray} \lambda_{h\tilde{t}_1\tilde{t}_1}^{m_{X}\gg M_Z} \ &\simeq& \ \frac{i g_W M_Z}{c_W} \biggl \{ \biggl [ -\frac{1}{2} \cos 2 \beta \cos^2\theta_{\tilde{t}} + \frac{2}{3} s_W^2 \cos 2\beta \cos 2\theta_{\tilde{t}} \biggr ] \nonumber \\[2mm] &-& \frac{m_t^2 } {M_Z^2 } -\frac{m_t \sin 2 \theta_{\tilde{t}}}{2 M_Z^2} \left ( A_t + \mu \cot \beta \right ) \biggr \} \;, \label{ht1t1aprox} \end{eqnarray} (here and in the following, $g_W^2=4\pi\alpha_{\mathrm{em}}/s^2_W$, $c_W=\sqrt{1-s^2_W}$ and $M_W^2=M_Z^2(1-s^2_W)$) where for large $A_t$ (note that $\sin\theta_{\tilde{t}}$ is maximal in such a case) the coupling goes like $\lambda_{h\tilde{t}_1\tilde{t}_1} \propto \frac {m_t A_t}{M_Z^2}$. As a consequence, because of the presence of the trilinear term $A_t$, the coupling of the light Higgs boson to light stop squarks could be much larger compared to that to the top quark, which behaves like $\lambda_{h t t} \propto \frac{m_t}{M_Z}$, as already recognised in \cite{djouadi,djouadi2}. Over the parameter space that we have chosen here, the cross section of $g g \rightarrow \tilde{t}_1 \tilde{t}_1^* h$ can be either larger than or of the same order as that of $g g \rightarrow t \bar{t} h$ \cite{djouadi,djouadi2}. Thus, the subprocess $gg\rightarrow \tilde{t}_1 \tilde{t}_1^* h$ can well boast the status of an additional discovery mechanism of the lightest scalar Higgs boson of the MSSM, as remarked in Refs.~\cite{djouadi,djouadi2}. (This is true also in non M-SUGRA models, where however one still has that $m_{H}\approx m_A\approx m_{H^\pm}\gg m_h$ \cite{djouadi,djouadi2}.) In this respect, the reader should further notice the stability of its production cross section against variations of $\tan\beta$ (see also, e.g., Fig. 4 in \cite{djouadi2})\footnote{The $\tan\beta$ dependence of $\sigma(\tilde{t}_1\tilde{t}_1^* h)$ at low values of such a parameter is mainly a phase space effect, as it can be deduced by comparing Fig.~\ref{fig:squarks} to Fig.~\ref{fig:lighthiggs}. In addition, in the low $\tan\beta$ domain, there are residual effects onto the production rates induced by the term $\mu \cot\beta$ arising from the off-diagonal elements of the squark mass matrices and affecting the $\lambda_{h {\tilde t}_1{\tilde t}_1}$ vertex.}. This proves to be a crucial point, as it is possible that one will have no narrow hints about the actual value of this crucial parameter of the Higgs sector even after Run 2 at Tevatron (unless, of course, the lightest Higgs boson is discovered there !) \cite{Tevatron}. In other terms, ${\tilde t}_1{\tilde t}_1^* h$ would always be present at fixed rate at the LHC, no matter whether $\tan\beta$ is large or small. Similar arguments can be put forward concerning the $\mathrm{sign}(\mu)$ dependence. Some care must instead be exercised with respect to the $A_0$ dependence. In fact, to vary the universal trilinear coupling between, e.g., $-300$ and $+300$ depletes the cross section by a factor of about seven, as shown in Fig.~\ref{fig:lighthiggs}. For even larger differences, say, between $-500$ and $+500$ (not shown here for reasons of space), the ratio between the cross sections become as big as 30 ! Not surprisingly then, Ref.~\cite{djouadi2} focused on the choice $A_0=-2000$ GeV\footnote{Note that in this $A_0$ region the M-SUGRA scenario clashes against the constraints from the charge and colour breaking minima, {i.e.}, eq.~(\ref{ccb}). This is the reason why we prefer to display our results in a rather more conservative range, i.e., $|A_0| \buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} 1$ TeV.} (and $\mathrm{sign}(\mu)=+$). Far from regarding this dependence as a shortcoming of ${\tilde t}_{1}{\tilde t}_{1}^*h$ events in helping in the quest for the so far elusive lightest Higgs boson of the MSSM (in fact, even for very large and positive $A_0$ values we found the cross section well above 1 fb), this example allows us to enlighten that other aspect of squark-squark-Higgs production that we have mentioned in the Introduction: i.e., its potential in pinning down some of the fundamental parameters of the M-SUGRA model. Needless to say, variations of the cross section with $A_0$ as large as those mentioned above are well beyond the various sources of uncertainties on the production rates (other than the theoretical ones related to the PDFs and the effect of higher-order QCD corrections also the experimental ones in their determination). To measure a production rate of ${\tilde t}_{1}{\tilde t}_{1}^*h$ events much larger than about 50 fb (the value obtained in correspondence of $A_0=0$), in some specific decay channel, would unambiguously imply that the universal trilinear coupling at the GUT scale is negative. As already mentioned, very little could be learn about the actual values of $\tan\beta$ from this specific process. However, if the latter is known beforehand to be around 2 or so, one could use this information to constrain $\mathrm{sign}(\mu)$. In fact, for $A_0=-300$ GeV, one would get that \begin{eqnarray} \sigma (\tilde{t}_1 \tilde{t}_1^* h) &\approx& 200 \; {\rm fb} \; \Rightarrow \mu = + \\[2mm] \sigma (\tilde{t}_1 \tilde{t}_1^* h) &\approx& 50 \; {\rm fb} \; \Rightarrow \mu = - \;. \end{eqnarray} Let us proceed in this spirit to see whether other channels can be of some help in constraining the M-SUGRA model. Following the list of detectable $h$ production cross section given in (\ref{Xsections_h}), we find the $\tilde{t}_1 \tilde{t}_2^* h$ final state \cite{short}. This is not surprising either. In fact, the relevant coupling behaves like (again, $X=H,A,H^\pm$) \begin{eqnarray} \lambda_{h\tilde{t}_1\tilde{t}_2}^{m_{X}\gg M_Z} \ &\simeq & \ \frac{i g_W M_Z}{c_W} \biggl \{ \frac{1}{2} \cos 2 \beta\biggl (\frac{1}{2}-\frac{4}{3}s_W^2 \biggr ) \sin2\theta_{\tilde{t}} \nonumber \\[2mm] &-&\frac{m_t \cos 2 \theta_{\tilde{t}}}{2 M_Z^2} \left ( A_t + \mu \cot \beta \right ) \biggr \} \;, \label{ht1t2aprox} \end{eqnarray} becoming very large when $\cos 2\theta_{\tilde{t}}$ and $ ( A_t + \mu \cot \beta ) $ reach their allowed maximum values. From Fig.~\ref{fig:lighthiggs}, one can see that this happens in the region of small $\tan\beta$, negative sign of $\mu$ and $A_0\buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} -300$ GeV. The intriguing aspect here, which was largely missing in the case in which both squarks were the lightest, is that one could impose severe constraints on the sign of the Higgsino mass term, other than on $A_0$. In fact, in the detectable region, the curves corresponding to $\mathrm{sign}(\mu)=-$ (higher) and $\mathrm{sign}(\mu)=+$ (lower) depart considerably. For example, for relatively small $\tan\beta$ values, say 4, the ratios as obtained by dividing the cross sections corresponding to negative $\mu$'s by those for positive Higgsino masses are quite large indeed: about 7(5)[2] when $A_0=-300(0)[+300]$ GeV. At even lower $\tan\beta$, say, equal to 2, one symbolically has: \begin{eqnarray} \sigma (\tilde{t}_1 \tilde{t}_2^* h) &\approx& 300 \; {\rm fb} \; \Rightarrow \mu = - \\[2mm] \sigma (\tilde{t}_1 \tilde{t}_2^* h) &\approx& 2 \; {\rm fb} \; \Rightarrow \mu = + \;, \end{eqnarray} (e.g., for $A_0=-300$ GeV). Luckily enough here, where the solid and dot-dashed curves start getting closer (for $\tan\beta\OOrd15-20$) is precisely when the cross section is no longer observable. However, what just said makes the point that $\tan\beta$ ought to be known rather accurately from some previous measurements, if one wants to constrain the other M-SUGRA parameters by studying the production of the lightest Higgs scalar of the theory produced in association with both stop mass eigenstates. In our list of observable $h$ cross sections ${\tilde{b}}_1{\tilde{b}}_1^* h$ comes next. Here the potential is somewhat complementary to the two above cases, in the sense that not to find any pairs of sbottom squarks of the type ${\tilde b}_1$ produced in association with an $h$ scalar once $\tan\beta$ is already known to be large could have powerful consequences on the viability of M-SUGRA as the underlying model of SUSY. To be specific, notice how the six curves corresponding to all the possible combinations of the parameters $A_0=-300,0,+300$ GeV and $\mathrm{sign}(\mu)=\pm$ lie within a factor from 2 to 4 in cross section, in correspondence of $\tan\beta=20$ and 35, respectively. Even the cases of $A_0=\pm500$ GeV do not depart significantly from the central curve for $A_0=0$, in the above $\tan\beta$ region. Unfortunately, contrary to the case of ${\tilde{t}}_1{\tilde{t}}_2^* h$ production, here the most interesting region is presumably below detection level. In fact, for $\tan\beta$ quite low, a huge portion of M-SUGRA parameter plane collapses into a narrow stripe, as the various curves tend to overlap, all being contained within a factor as small as 1.5 (e.g., at $\tan\beta=2$, also including the cases $|A_0|=500$ GeV, not shown in the figure). As for $H$ production, one identifies as possible candidates for detection the following cases: \begin{equation}\label{Xsections_H} \sigma( \tilde{b}_1 \tilde{b}_1^* H )\sim \sigma( \tilde{t}_1 \tilde{t}_1^* H )\buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} \sigma( \tilde{b}_2 \tilde{b}_2^* H )\sim \sigma( \tilde{t}_1 \tilde{t}_2^* H ). \end{equation} The remaining two combinations, i.e., $\tilde{b}_1 \tilde{b}_2^* H$ and $\tilde{t}_2 \tilde{t}_2^* H$, yield cross sections that are hopelessly small. Also some of the detectable $H$ production processes can have a significant impact in aiding the determination of the M-SUGRA parameters, most notably those yielding the final states ${\tilde{b}}_1{\tilde{b}}_1^* H$ and ${\tilde{b}}_2{\tilde{b}}_2^* H$. Here, if $\tan\beta$ is known to be, say, 35 or so, a detection of either the former or the latter by the thousand or hundred, respectively, would imply that $A_0$ is most certainly negative, since production rates corresponding to $A_0$ values larger than zero are about a factor of 5 and 3 smaller (rather irrespectively of $\mathrm{sign}(\mu)$). Somewhat less discrimination power between positive and negative $A_0$ values have ${\tilde{t}}_1{\tilde{t}}_1^* H$ and ${\tilde{t}}_1{\tilde{t}}_2^* H$ events, over the same (large) $\tan\beta$ region as above. The most interesting case would have been ${\tilde{t}}_2{\tilde{t}}_2^* H$, as the collapse of the M-SUGRA model that we already noticed in the case of ${\tilde{b}}_1{\tilde{b}}_1^* h$ final states is here even more striking, over a more significant $\tan\beta$ region. Unluckily enough, the corresponding production cross section never exceeds the femtobarn level. A general remark in now in order, concerning the strength of the coupling of sbottom squarks to neutral CP-even Higgs bosons. The monotonic growth of the production rates of sbottom squark processes with increasing $\tan\beta$, as opposed to a much milder dependence of the stop ones, has a simple explanation. For example, the coupling $\lambda_{h \tilde{b}_1 \tilde{b}_1}$ is driven by the term $\frac{m_b \mu \tan\beta}{M_Z^2}$ and its size becomes large for large $\tan\beta$. Another reason for an extra enhancement of the sbottom production rates comes from the phase space available to the final states, as both $m_{\tilde{b}_1}$ and $ m_{\tilde{b}_2}$ decrease very fast when $\tan\beta$ gets large, whereas this is much less the case for $m_{\tilde{t}_1}$ and $ m_{\tilde{t}_2}$ (see Fig.~\ref{fig:squarks}). There is however a point, for the case of reactions involving the two sbottom mass eigenstates at once (i.e., ${\tilde{b}_1}{\tilde{b}_2}^*h$ and ${\tilde{b}_1}{\tilde{b}_2}^*H$), in which the production cross sections vanish altogether, somewhere in the vicinity of $\tan\beta=34-36$ (the zero for ${\tilde{b}_1}{\tilde{b}_2}^*H$ is actually beyond the $\tan\beta$ range plotted), the exact value depending upon $A_0$ and $\mathrm{sign}(\mu)$. This is clearly induced by the $\lambda_{h {\tilde{b}}_1{\tilde{b}}_2}$ and $\lambda_{H {\tilde{b}}_1{\tilde{b}}_2}$ vertices and their typical $\propto(\mu-A_b \tan\beta)$ behaviour, when $|\mu|\ll | A_b \tan\beta|$ and $A_b$ changes its sign. Before closing this Section, we investigate the residual dependence of CP-even Higgs boson production in association with sbottoms and stops on the input values of $M_{0}$ and $M_{1/2}$, when they are allowed to deviate from their common default of 150 GeV assumed so far. For illustrative purposes, we do so by adopting two discrete values of $\tan\beta$, 2 and 40, and choosing the combination $A_0=0$ and negative $\mu$. Anyhow, though not shown, we have verified that a similar pattern to the one that we will outline below can be recognised also for the case of finite (positive and negative) values of $A_0$ and positive Higgsino masses as well. For this exercise, we focus our attention only to the dominant production cross sections in either case, that is, $\tilde{t}_1 \tilde{t}_1^* h$, $\tilde{t}_1 \tilde{t}_2^* h$ and $\tilde{b}_1 \tilde{b}_1^* h$ for light (see Tab.~\ref{table1}) and $\tilde{t}_1 \tilde{t}_1^* H$ and $\tilde{b}_1 \tilde{b}_1^* H$ for heavy (see Tab.~\ref{table2}) Higgs bosons. \begin{table} \begin{eqnarray} \begin{array}{|c|c|c|c|c|c|} \hline M_0 ({\rm GeV}) & M_{1/2} ({\rm GeV}) & \tan\beta & \sigma (g g \rightarrow \tilde{t}_1 \tilde{t}_1^* h ) ({\rm fb}) & \sigma (g g \rightarrow \tilde{t}_1 \tilde{t}_2^* h ) ({\rm fb}) & \sigma (g g \rightarrow \tilde{b}_1 \tilde{b}_1^* h ) ({\rm fb}) \\ \hline 130 & 130 & 2 & 70.2 & 38 & 7.7 \times 10^{-2} \\ \hline 200 & 150 & 2 & 32 & 150 & 2.9 \times 10^{-2} \\ \hline 200 & 200 & 2 & 11 & 100 & 6.6 \times 10^{-3} \\ \hline 300 & 250 & 2 & 2.9 &48 & 1.4\times 10^{-3} \\ \hline 130 & 130 & 40 & 84 & 8.2 & 8.2 \\ \hline 200 & 150 & 40 & 37 & 7.4 & 4.9 \\ \hline 200 & 200 & 40 & 13 & 3.4 & 1.3 \\ \hline 300 & 250 & 40 & 3.5 & 1.8 & 0.51 \\ \hline \end{array} \nonumber \end{eqnarray} \caption{The variation of the most significant cross sections of processes $g g \rightarrow \tilde{q}_\chi \tilde{q}_{\chi '}^* h $ with $M_0$, $M_{1/2}$ and $\tan\beta$. The other M-SUGRA parameters are fixed as follows: $A_0=0$ GeV and $\mathrm{sign}(\mu) = -$.} \label{table1} \end{table} We obtain that most of the cross sections with the light Higgs particle involved decrease when either or both the parameters $M_0$ and $M_{1/2}$ increase. In this respect, however, it is well worth noticing that the total cross section for $\tilde{t}_1 \tilde{t}_2^* h $ production acquires a large statistic significance in the higher mass regime and maintains it even at the very upper end of it (some 5000 events/year can be produced at the LHC via this mode if $\tan\beta=2$, $M_0=300$ GeV and $M_{1/2}=250$ GeV). In this area of the M-SUGRA parameter space, $\tilde{t}_1 \tilde{t}_2^* h $ events are much more numerous than $\tilde{t}_1 \tilde{t}_1^* h $ ones, the other way round with respect to the low mass combination $M_0=M_{1/2}=130$ GeV, despite of the more massive final state. (However, this is only true al low $\tan\beta$.) In fact, for $M_0=300$ GeV and $M_{1/2}=250$ GeV, the squark masses are $m_{\tilde{t}_1}=472$ GeV, $m_{\tilde{t}_2}=591$ GeV, $m_{\tilde{b}_1}=569$ GeV and $m_{\tilde{b}_2}=623$ GeV. The inversion of hierarchy between the two cross sections is induced by the onset of the ${\tilde{t}_2}\to {\tilde{t}_1} h$ decay channel at large $M_{0}$ and $M_{1/2}$ values, as $m_{\tilde{t}_2}\buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} m_{\tilde{t}_1} +m_h$, whose resonance enhancement in the $2\to2$ process $gg\to {\tilde{t}}_2{\tilde{t}}_2^*$ overcomes both the inner phase space depletion and the strength of the Higgs-strahlung emission in $gg\to {\tilde{t}}_1{\tilde{t}}_1^*$ events. As for the case $\tilde{b}_1 \tilde{b}_1^* h $ (and similarly for $\tilde{b}_2 \tilde{b}_2^* h $, not shown in the table), there is no inversion of tendency here, in the interplay with the light stop channel, as production rates are strongly dominated by the fact that $m_{\tilde{b}_2}, m_{\tilde{b}_1} \gg m_{\tilde{t}_1}$. They both are very much suppressed. On similar grounds, one can argue about the smallness of $\tilde{t}_2 \tilde{t}_2^* h $. Finally, being $m_{\tilde{b}_2}\buildrel{\scriptscriptstyle <}\over{\scriptscriptstyle\sim} m_{\tilde{b}_1} +m_h$ in most part of the ($M_0$,$M_{1/2}$) plane that we have spanned, the $\tilde{b}_2 \tilde{b}_1^* h $ final state never stands up either as quantitatively interesting. The results for the mass dependence of the production rates for the heavy (CP-even) Higgs boson have a simpler pattern. The all of their production phenomenology is governed by the fact that over the ($M_0$,$M_{1/2}$) regions considered here one never finds the kinematic configuration $m_{\tilde{t(b)}_2}\buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} m_{\tilde{t(b)}_1} +m_H$. That is, no production and decay channel can onset, and the hierarchy of cross sections already seen for $M_0=M_{1/2}=150$ GeV replicates unaltered in most cases, mainly governed by the size of the rest masses in the final state. Here cross sections remain sizable only if neither $M_0$ nor $M_{1/2}$ exceed the value 150--200 GeV (for large $\tan\beta$, of course, see Fig.~\ref{fig:heavyhiggs}). Even in such cases, though, presumably no more than a handful of events can be selected in most decay channels. \begin{table} \begin{eqnarray} \begin{array}{|c|c|c|c|c|} \hline M_0 ({\rm GeV}) & M_{1/2} ({\rm GeV}) & \tan\beta & \sigma (g g \rightarrow \tilde{t}_1 \tilde{t}_1^* H ) ({\rm fb}) & \sigma (g g \rightarrow \tilde{b}_1 \tilde{b}_1^* H ) ({\rm fb}) \\ \hline 130 & 130 & 2 & 8.0 \times 10^{-3} & 2.1\times 10^{-3} \\ \hline 200 & 150 & 2 & 1.4 \times 10^{-3} & 4.2\times 10^{-4} \\ \hline 200 & 200 & 2 & 1.1 \times 10^{-4} & 7.5\times 10^{-5} \\ \hline 300 & 250 & 2 & 1.3 \times 10^{-5} & 8.2\times 10^{-6} \\ \hline 130 & 130 & 40 & 3.3 & 14 \\ \hline 200 & 150 & 40 & 1.3 & 3.5 \\ \hline 200 & 200 & 40 & 0.38 & 0.84 \\ \hline 300 & 250 & 40 & 0.068 & 0.014 \\ \hline \end{array} \nonumber \end{eqnarray} \caption{The variation of the most significant cross sections (in pb) of processes $g g \rightarrow \tilde{q}_\chi \tilde{q}_{\chi '}^* H $ with $M_0$, $M_{1/2}$ and $\tan\beta$. The other M-SUGRA parameters are fixed as follows: $A_0=0$ GeV and $\mathrm{sign}(\mu) = -$.} \label{table2} \end{table} \subsubsection{CP-odd Higgs boson production} \label{odd} As discussed to some lenght in Ref.~\cite{short}, pseudoscalar Higgs boson production in association with sbottom and stop squarks of different mass (the only possible combination in absence of CP-violating phases in $\mu$ and $A_q$, with $q=t,b$), can boast a special attractiveness because of the absence of mixing terms in the relevant squark-squark-Higgs couplings. By making use of eq.~(\ref{fr}) and recalling that if one reverts the chirality flow in the vertex $\lambda_{A\tilde{q}_L \tilde{q}_R}$ the corresponding Feynman rule changes its sign \cite{HHG}, one finds that those vertices reduce to \begin{equation} \lambda_{A\tilde{t}_1\tilde{t}_2} = -\frac{g_W m_t}{2 M_W} \left ( \mu - A_t \cot\beta \right), \qquad\qquad \lambda_{A\tilde{b}_1\tilde{b}_2} = -\frac{g_W m_b}{2 M_W} \left ( \mu - A_b \tan\beta \right ). \label{cpoddfr} \end{equation} These are precisely the couplings entering the two processes of the type (\ref{proc}) inducing the final states ${\tilde{q}}_1{\tilde{q}}_2^* A$, where $q=t,b$. From this point of view, it is then clear the potential of squark and pseudoscalar Higgs production in constraining the input values of all five M-SUGRA parameters. In other terms, to trace back (more technically, to fit) the shape of the cross sections (if not of some differential distributions) in terms of the $\tan\beta$, $A_q$ (with $q=t,b$) and $\mu$ parameters entering eq.~(\ref{cpoddfr}) is presumably a much simpler job then doing the same by using the more involved expressions in eq.~(\ref{fr}), unless one exploits some asymptotic regime in either $\tan\beta$, $A_q$ (with $q=t,b$) and/or $\mu$ in which the latter reduce to the former. It is under this perspective that we looked at the case of $A$ production in our Ref.~\cite{short}. Rather than repeating the all discussion carried out there, we summarise here the salient findings of \cite{short}, referring the reader to that paper for specific details. The production cross sections can be found in Fig.~\ref{fig:cpoddhiggs}. For $\tan\beta$ below 20 or so, the rates for pseudoscalar Higgs boson production are presumably too poor to be of great experimental help. Furthermore, in the high $\tan\beta$ regime, pseudoscalar Higgs boson production is in general less effective than other channels in constraining the sign of the Higgsino mass term: compare the overlapping for the solid and dot-dashed curves (for each $A_0$) in the detectable regions of $\tilde{t}_1 \tilde{t}_2^* A$ and $\tilde{b}_1 \tilde{b}_2^* A$ production to the splitting occurring in, e.g., $\tilde{t}_1 \tilde{t}_1^* H$. This, as far as it concerns the flaws. As for the advantages, we would like to draw the attention of the reader to the fact that reactions (\ref{proc}) with CP-odd Higgs bosons in the final state are quite sensitive to $\tan\beta$. The simple form of the expressions for $\lambda_{A {\tilde q}_1{\tilde q}_2}$ ($q=t,b$) in eq.~(\ref{cpoddfr}) allows one to straightforwardly interpret the variation of the pseudoscalar rates with this parameter, namely, the steep rise at high values of the latter. This can in fact be understood as follows. For large $\tan\beta$, the vertex couplings of eq.~(\ref{cpoddfr}) can be rewritten in the approximate form \begin{equation}\label{limit} \lambda_{A\tilde{t}_1\tilde{t}_2} \simeq -\frac{g_W m_t}{2 M_W} \mu , \qquad\qquad \lambda_{A\tilde{b}_1\tilde{b}_2} \simeq \frac{g_W m_b}{2 M_W} A_b \tan\beta , \label{largetanb} \end{equation} that is, the coupling which is associated with the sbottom pair is proportional to $\tan\beta$, so that, eventually, the total $\tilde{b}_1\tilde{b}_2^*A$ cross section will grow with $\tan^2\beta$ while the coupling related to the stop pair will assume constant values. In the latter, the enhancement of the $\tilde{t}_1\tilde{t}_2^*A$ cross section with $\tan\beta$ is rather a phase space effect since, as $\tan\beta$ increases, the CP-odd Higgs boson mass decreases considerably (the squark masses changing much less instead), as we can see from Fig.~\ref{fig:higgs}. Of course, the same remains valid in the former case as well, so that our figure indicates a clear order in the size of the cross sections, $\sigma (\tilde{b}_1 \tilde{b}_2^* A ) \buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} \sigma (\tilde{t}_1 \tilde{t}_2^* A )$, at large $\tan\beta$. But, let us now turn our attention to another peculiar dependence of the production rates of $\tilde{t}_1 \tilde{t}_2^* A$ and $\tilde{b}_1 \tilde{b}_2^* A$: the one on the common trilinear coupling $A_0$. Pretty much along the same lines as for the combinations $\tilde{t}_1 \tilde{t}_1^* h$, $\tilde{t}_1 \tilde{t}_2^* h$ and $\tilde{b}_1 \tilde{b}_1^* H$ one can make the case that the sensitivity to $A_0$ of the $A$ production cross sections offers the chance of constraining, possibly the sign, and hopefully the magnitude, of this fundamental M-SUGRA parameter. This is presumably the best attribute of $\tilde{t}_1 \tilde{t}_2^* A$ and $\tilde{b}_1 \tilde{b}_2^* A$, under the assumption already made in few instances that the determination of $\tan\beta$ could come first from studies in the pure Higgs sector. Putting down some numbers in this respect, one may invoke the following scenario, if $\tan\beta$ is, say, larger than 32: \begin{eqnarray} \sigma (\tilde{t}_1 \tilde{t}_2^* A) &\buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim} & 10 \; {\rm fb} \; \Rightarrow A_0 < -300 \\[2mm] \sigma (\tilde{b}_1 \tilde{b}_2^* A) &\buildrel{\scriptscriptstyle >}\over{\scriptscriptstyle\sim}& 2 \; {\rm fb} \; \Rightarrow A_0 < -300, \end{eqnarray} quite independently of $\mathrm{sign}(\mu)$. Conversely, the non-observation of pseudoscalar Higgs events in those regimes would imply most likely a positive $A_0$ value. As for peculiar trends in Fig.~\ref{fig:cpoddhiggs}, it is worth mentioning (though we have not shown it here, as the reader can refer to \cite{short}) that the cross section for sbottom production vanishes too, at some (large) value of $\tan\beta$, as it did for the case of CP-even Higgs production. As the reader can appreciate in Fig.~2 of Ref.~\cite{short}, contrary to the case of $\tilde{b}_1 \tilde{b}_2^* h$ and $\tilde{b}_1 \tilde{b}_2^* H$ final states, this however happens in $\tilde{b}_1 \tilde{b}_2^* A$ events only for positive and large values of $A_0$ (and both $\mathrm{sign}(\mu)=\pm$). This is another consequence of the different nature of the couplings (\ref{fr}) to squarks of CP-even versus CP-odd neutral Higgs bosons. Though we have failed to find a point where this disappearance of $\tilde{b}_1 \tilde{b}_2^* A$ events for positive $A_0$ values corresponds to the survival of a detectable cross section for negative $A_0$'s (at fixed $\tan\beta$), such matter would presumably deserve further investigation in the future. Before closing this Section, we study the dependence of pseudoscalar Higgs boson production in association with stop and sbottom squarks on the last two M-SUGRA independent parameters, $M_0$ and $M_{1/2}$: see Tab.~\ref{table3}. The main effect of changing the latter is onto the masses of the final state scalars, through the phase space volume as well as via propagator effects in the scattering amplitudes. (In fact, no decay channel of the heavier stop or sbottom into the lighter one ever opens, at least for the values of $M_0$ and $M_{1/2}$ that we had looked at.) In other terms, to increase one or the other depletes both $\sigma (\tilde{t}_1 \tilde{t}_2^* A )$ and $\sigma (\tilde{b}_1 \tilde{b}_2^* A )$ quite strongly, simply because the values of all $m_{\tilde{q}_\chi}$'s, $m_{\tilde{q}_{\chi'}}$'s and $m_\Phi$'s get larger. For example, assuming $\tan\beta=40$: at $M_0=M_{1/2}=130$ GeV, one has $m_{\tilde{t}_1}=248$ GeV, $m_{\tilde{t}_2}=388$ GeV, $m_{\tilde{b}_1}=256$ GeV, $m_{\tilde{b}_2}=340$ GeV and $m_{{A} }=120$; whereas at $M_0=300$ GeV and $M_{1/2}=250$ GeV, the figures are $m_{\tilde{t}_1}=461$ GeV, $m_{\tilde{t}_2}=611$ GeV, $m_{\tilde{b}_1}=510$ GeV, $m_{\tilde{b}_2}=591$ GeV and $m_{{A} }=292$ GeV. In practice, the table shows that only rather light $M_0$ and $M_{1/2}$ masses (say, below 200 and 150 GeV, respectively) would possibly allow for pseudoscalar production to be detectable at the LHC, and only at large $\tan\beta$ \cite{short}. \begin{table} \begin{eqnarray} \begin{array}{|c|c|c|c|c|} \hline M_0 ({\rm GeV}) & M_{1/2} ({\rm GeV}) & \tan\beta & \sigma (g g \rightarrow \tilde{t}_1 \tilde{t}_2^* A ) ({\rm fb}) & \sigma (g g \rightarrow \tilde{b}_1 \tilde{b}_2^* A ) ({\rm fb}) \\ \hline 130 & 130 & 2 & 5.2 \times 10^{-2} & 4.1\times 10^{-4} \\ \hline 200 & 150 & 2 & 2.0 \times 10^{-2} & 8.8\times 10^{-5} \\ \hline 200 & 200 & 2 & 5.9 \times 10^{-3} & 2.8\times 10^{-5} \\ \hline 300 & 250 & 2 & 1.3 \times 10^{-3} & 3.9\times 10^{-6} \\ \hline 130 & 130 & 40 & 79 & 13 \\ \hline 200 & 150 & 40 & 1.4 & 2.4 \\ \hline 200 & 200 & 40 & 0.31 & 0.6 \\ \hline 300 & 250 & 40 & 0.048 & 0.098 \\ \hline \end{array} \nonumber \end{eqnarray} \caption{The variation of the most significant cross sections of processes $g g \rightarrow \tilde{q}_\chi \tilde{q}_{\chi '}^* A $ with $M_0$, $M_{1/2}$ and $\tan\beta$. The other M-SUGRA parameters are fixed as follows: $A_0=0$ GeV and $\mathrm{sign}(\mu) = -$.} \label{table3} \end{table} \subsubsection{Charged Higgs bosons production} \label{charged} To have an additional source of charged Higgs bosons at the LHC, especially with masses larger than the top mass $m_t$, would be very helpful from an experimental point of view. In fact, it is well known the difficulty of detecting charged Higgs scalars in that mass regime, not only because of a dominant decay signature which suffers from very large QCD background (i.e., $H^+\to t\bar b \to b\bar b W^+\to b\bar b jj$), but also because the production mechanisms are not many and with not very large rates \cite{charged}. Unfortunately, as it turns out from Fig.~\ref{fig:chargedhiggs}, typical production cross sections of $H^\pm$ scalars in association with sbottom and stop pairs rarely exceed 10 fb. These rates compare rather poorly with other mechanisms \cite{charged}, for the same choice of $m_{H^\pm}$. Thus, there is little to gain in exploiting processes (\ref{proc}) as discovery channels of charged Higgs bosons. Furthermore, their dependence on $\tan\beta$, $A_0$ and $\mathrm{sign}(\mu)$ replicates many of the tendencies already individuated in neutral Higgs channels, for which the production cross sections are much larger. Adding the fact that typical decay channels of the latter (e.g., in photon pairs) are much cleaner in the hadronic environment of the LHC than those of the former, one would quite rightly conclude that the potential of $\tilde{t}_1 \tilde{b}_1^* H^-$, $\tilde{t}_1 \tilde{b}_2^* H^-$, $\tilde{b}_1 \tilde{t}_2^* H^+$ and $\tilde{b}_2 \tilde{t}_2^* H^+$ final states in constraining the M-SUGRA parameter space is rather poor. Nonetheless, it is worth recognising some of the typical trends of the production cross sections, for the sake of future reference. Let alone the last two combinations, for which the final state masses are too heavy to be produced at detectable rate, we have a quick look at the first two cases, which can indeed have cross sections significantly above 1 fb, at least in the large $\tan\beta$ region. This enhancement has a twofold explanation. Firstly, phase space effects, as for large $\tan\beta$ all scalar masses (except $m_{{\tilde t}_1}$ and $m_h$) get smaller: see Figs.~\ref{fig:squarks}--\ref{fig:higgs}. Secondly, terms in their couplings proportional to $m_b A_b \tan\beta$ are dominant for most of the possible $A_0$ and $\mathrm{sign}(\mu)$ combinations. Finally, notice also in the case of the $\tilde{t}_1 \tilde{b}_1^* H^\pm$ final state the vanishing of the cross section, this time at somewhat lower values of $\tan\beta$ than in the case of the neutral Higgs bosons. As for the $M_0$ and $M_{1/2}$ dependence, this is again realised through the phase space and the propagators, as there is no significant enhancement from resonant decays. In practice, only if $\tan\beta$ is extremely large and both the universal scalar and gaugino masses are below 200 GeV, the two cross sections for $\tilde{t}_1 \tilde{b}_1^* H^-$ and $\tilde{t}_1 \tilde{b}_2^* H^-$ remain detectable (indeed, those containing the lightest stop squark): see Tab.~\ref{table4}}. \begin{table} \begin{eqnarray} \begin{array}{|c|c|c|c|c|} \hline M_0 ({\rm GeV}) & M_{1/2} ({\rm GeV}) & \tan\beta & \sigma (g g \rightarrow \tilde{t}_1 \tilde{b}_1^* H^- ) ({\rm fb}) & \sigma (g g \rightarrow \tilde{t}_1 \tilde{b}_2^* H^- ) ({\rm fb}) \\ \hline 130 & 130 & 2 & 4.3 \times 10^{-4} & 1.4\times 10^{-3} \\ \hline 200 & 150 & 2 & 2.2 \times 10^{-5} & 3.0\times 10^{-4} \\ \hline 200 & 200 & 2 & 3.1 \times 10^{-5} & 9.8\times 10^{-5} \\ \hline 300 & 250 & 2 & 1.6 \times 10^{-6} & 1.4\times 10^{-5} \\ \hline 130 & 130 & 40 & 2.2 & 4.8 \\ \hline 200 & 150 & 40 & 1.5 & 4.5 \\ \hline 200 & 200 & 40 & 1.7 & 0.31 \\ \hline 300 & 250 & 40 & 0.064 & 0.023 \\ \hline \end{array} \nonumber \end{eqnarray} \caption{The variation of the most significant cross sections of processes $g g \rightarrow \tilde{q}_\chi \tilde{q}_{\chi '}^* H^\pm $ with $M_0$, $M_{1/2}$ and $\tan\beta$. The other M-SUGRA parameters are fixed as follows: $A_0=0$ GeV and $\mathrm{sign}(\mu) = -$.} \label{table4} \end{table} \subsubsection{Decay signatures} \label{signatures} So far we have only discussed production cross sections for processes of the form (\ref{proc}) and made no considerations about possible decay channels and relative branching fractions of either squarks or Higgs bosons. Another related aspect is the typical kinematics of the signals, as it appears in the detectors, and the size of the possible backgrounds. Furthermore, the reader should appreciate how all channels entering processes (\ref{proc}) are intertwined, in the sense that any of these can act as a background to all others. It is the aim of this Section that of indicating some possible decay signatures of the most relevant squark-squark-Higgs processes, in which they show both large rates and their kinematics is such that they can hopefully be disentangled at the LHC. In doing so, we distinguish between a small (see \ref{signatures}.1) and large (see \ref{signatures}.2) $\tan\beta$ regime, as we have shown that such a parameter is crucial in determining the actual size of the production rates. As representative choice of the universal masses we adopt the combination with lowest values among those discussed in the previous Subsections, i.e., $M_{0}=M_{1/2}=130$ GeV, further setting $A_0$=0 and $\mathrm{sign}(\mu)$ negative. \vskip0.5cm\noindent {\bf \ref{signatures}.1 Small $\tan\beta$ regime} \vskip0.5cm\noindent For small $\tan\beta$'s the only relevant processes are $\tilde{t}_1 \tilde{t}_1^* h$ and $\tilde{t}_1 \tilde{t}_2^* h$ production. A possible decay signature for the first case is the one contemplated in Refs.~\cite{djouadi,djouadi2}. That is, ${\tilde t}_1\to \chi_1^+b\to W^{+}b$ plus missing energy for the light stop and $h\to \gamma\gamma$ for the light Higgs boson, with the $W^+$ decaying leptonically and/or hadronically. The final topology would be the same as in $t\bar t h$, with the only difference that for stop squark events there is a large amount of missing energy. Since another option to tag the lightest MSSM Higgs boson at the LHC is to use the more messy but dominant decay channel into $b\bar b$ pairs (as opposed to exploiting the cleaner but suppressed $\gamma\gamma$ mode) \cite{ATLAS,CMS}, another possible decay sequence could be the following: $$ \arraycolsep=0pt \begin{array}{lllllll}\label{hchain} {\tilde{t}_1} && &~~~~{\tilde{t}}_1^* &&h~~~~~ & \\ \downarrow &&& ~~~\downarrow &&\downarrow~~~~~ & \\ \chi_1^+ \,+\,b\; & &&~~~~\chi_1^- \,+\,\bar b &~&b \,+\,\bar b & \\ \downarrow &&& ~~~\downarrow && \; & \\ q \,+\,\bar q'\,+\,\chi_1^0& && ~~~~\ell^-\,+\,\nu\,+\,\chi^0_1 &&& \end{array} $$ in which $q\bar q'=u\bar d,c\bar s$ and $\ell=e,\mu$. Considering also the charge conjugated $\chi_1^+\chi_1^-$ decays, the final signature would then be `$2~{\mathrm{jets}}~+~4b~+\ell^\pm+{E}_{\mathrm{miss}}$', where the missing energy/momentum is not only due to the two $\chi^0_1$'s but also to the neutrinos. The total branching ratio (BR) of such a decay sequence is, for $\tan\beta=3$ and according to Tab.~\ref{table5}, approximately 2.5\%. The production cross section at the same $\tan\beta$ value is about 72 fb, so that about 176 events per year would survive. One may further assume a reduction factor of about 0.25 because of the overall efficiency $\varepsilon_b^4$ to tag four $b$-quarks (assuming $\varepsilon_b\approx0.7$). This ultimately yields something more than 44 events per year. In addition, one should expect most of the signal events to lie in the detector acceptance region, since leptons and jets originate from decays of heavy objects. The same signature could well be exploited in the case of the ${\tilde{t}_1} {\tilde{t}}_2^* h$ final state. Here, the production cross section is smaller than for ${\tilde{t}_1} {\tilde{t}}_1^* h$ production, about 40 fb, so is the ${\tilde t}_2\to \chi_1^+b$ decay rate as compared to the ${\tilde t}_1\to \chi_1^+b$ one (see Tab.~\ref{table5}). However, the final number of events per year is still quite large: about 14 after having already multiplied by $\varepsilon_b^4$. As for the kinematics of these two signatures, we may remark that they have peculiar features that should help in their selection: a not too large hadronic multiplicity, six jets in total, each rather energetic (in fact, note that $m_{\chi^\pm_1}-m_{\chi^0_1}\approx58$ GeV and $m_{{\tilde{t}_2}}\approx378~{\mathrm{GeV}}\gg m_{{\tilde{t}_1}}\approx273~{\mathrm{GeV}}\gg m_{\chi^\pm_1}\approx114$ GeV), so that their reconstruction from the detected tracks should be reasonably accurate; high transverse momentum and isolated leptons to be used as trigger; large $E_{\mathrm{miss}}$ to reduce non-SUSY processes; four tagged $b$-jets that can be exploited to suppress the `$W^\pm$ + light jet' background from QCD, and one $b\bar b$ pair resonating at the $h$ mass, $m_h\approx90$ GeV. Moreover, the `irreducible' background from ${\tilde{t}_1}{\tilde{t}_1^*}Z$ events has been shown in Ref.~\cite{fawzi} to be under control, even when $m_h\approx M_Z$, as it is the case here. \begin{center} \begin{table} \begin{eqnarray} \begin{array}{cc} \begin{array}{|ccc|}\hline {\rm Particle} & {\mathrm{BR}} & {\rm Decay} \\ \hline \tilde{t}_1 & \stackrel{76\% }{\rightarrow}& \chi_1^+ b \\ &\stackrel{19\% }{\rightarrow}& \chi_1^0 t \\ \hline \tilde{t}_2 & \stackrel{57\% }{\rightarrow}& \chi_1^+ b \\ & \stackrel{24\% }{\rightarrow} & \chi_2^+ b \\ & \stackrel{11\% }{\rightarrow} & \chi_2^0 t \\ \hline h & \stackrel{90\% }{\rightarrow}& b \bar{b} \\ & \stackrel{5\% }{\rightarrow} & \tau^+ \tau^- \\ & \stackrel{0.0003\% }{\rightarrow} & \gamma \gamma \\ \hline \end{array} & \begin{array}{|ccc|}\hline {\rm Particle} & {\mathrm{BR}} & {\rm Decay} \\ \hline t & \stackrel{33\% }{\rightarrow} & u \bar{d} b \\ & \stackrel{33\% }{\rightarrow} & c \bar{s} b \\ & \stackrel{11\% }{\rightarrow} & e^+ \nu b \\ & \stackrel{11\% }{\rightarrow} & \mu^+ \nu b \\ & \stackrel{11\% }{\rightarrow} & \tau^+ \nu b \\ \hline \chi_1^+ & \stackrel{30\% }{\rightarrow} & \chi_1^0 u \bar{d}^{\dagger} \\ & \stackrel{30\% }{\rightarrow} & \chi_1^0 c \bar{s}^{\dagger} \\ & \stackrel{14\% }{\rightarrow} & \chi_1^0 \tau^+ \nu^{\dagger} \\ & \stackrel{14\% }{\rightarrow} & \chi_1^0 e^+ \nu^{\dagger} \\ & \stackrel{14\% }{\rightarrow} & \chi_1^0 \mu^+ \nu^{\dagger} \\ \hline \chi_2^0 & \stackrel{28\% }{\rightarrow} & \chi_1^0 \nu \bar{\nu} \\ & \stackrel{14\% }{\rightarrow} & \chi_1^0 e^- e^+ \\ & \stackrel{14\% }{\rightarrow} & \chi_1^0 \mu^- \mu^+ \\ & \stackrel{14\% }{\rightarrow} & \chi_1^0 \tau^- \tau^+ \\ \hline \end{array} \end{array} \nonumber \end{eqnarray} \caption{Dominant decay channels and BRs of the final state (s)particles in $g g \rightarrow \tilde{q}_\chi \tilde{q}_{\chi '}^* h $, $q=t$ and $\chi,\chi'=1,2$, for $M_0=M_{1/2}=130$ GeV, $A_0=0$, $\tan\beta=3$ and ${\mathrm{sign}}(\mu)<0$ \cite{isajet}. {~~$^\dagger$ \footnotesize{Via off-shell $W^{+}$.}}} \label{table5} \end{table} \end{center} \begin{center} \begin{table} \begin{eqnarray} \begin{array}{cc} \begin{array}{|ccc|}\hline {\rm Particle} & {\mathrm{BR}} & {\rm Decay} \\ \hline \tilde{t}_1 & \stackrel{94\% }{\rightarrow}& \chi_1^+ b \\ \hline \tilde{t}_2 & \stackrel{40\% }{\rightarrow}& \chi_2^+ b \\ & \stackrel{26\% }{\rightarrow} & \chi_1^+ b \\ & \stackrel{16\% }{\rightarrow} & {\tilde{b}}_1 W^+ \\ & \stackrel{7\% }{\rightarrow} & {\tilde{t}}_1 Z \\ \hline \tilde{b}_1 & \stackrel{61\% }{\rightarrow}& \chi_2^0 b \\ & \stackrel{32\% }{\rightarrow} & \chi_1^0 b \\ \hline \tilde{b}_2 & \stackrel{42\% }{\rightarrow}& \chi_3^0 b \\ & \stackrel{31\% }{\rightarrow} & \chi_4^0 b \\ & \stackrel{18\% }{\rightarrow} & \chi_2^0 b \\ \hline h & \stackrel{94\% }{\rightarrow}& b \bar{b} \\ & \stackrel{6\% }{\rightarrow} & \tau^+ \tau^- \\ \hline H & \stackrel{94\% }{\rightarrow}& b \bar{b} \\ & \stackrel{6\% }{\rightarrow} & \tau^+ \tau^- \\ \hline A & \stackrel{94\% }{\rightarrow}& b \bar{b} \\ & \stackrel{6\% }{\rightarrow} & \tau^+ \tau^- \\ \hline H^\pm & \stackrel{91\% }{\rightarrow} & \tau^\pm \nu \\ & \stackrel{5\% }{\rightarrow} & \chi_1^0 \chi_1^\pm \\ \hline \end{array} & \begin{array}{|ccc|}\hline {\rm Particle} & {\mathrm{BR}} & {\rm Decay} \\ \hline \chi_1^+ & \stackrel{100\% }{\rightarrow} & \tilde{\tau}_1^+ \nu \\ \hline \chi_2^+ & \stackrel{24\% }{\rightarrow} & \chi_2^0 W^+ \\ & \stackrel{15\% }{\rightarrow} & \chi_1^+ Z \\ & \stackrel{11\% }{\rightarrow} & \chi_1^+ A \\ & \stackrel{10\% }{\rightarrow} & \tau^+ \tilde{\nu} \\ \hline \chi_2^0 & \stackrel{100\% }{\rightarrow} & \tilde{\tau}_1^\pm \tau^\mp \\ \hline \chi_3^0 & \stackrel{24\% }{\rightarrow} & \chi_1^+ W^- \\ & \stackrel{24\% }{\rightarrow} & \chi_1^- W^+ \\ & \stackrel{9\% }{\rightarrow} & \chi_1^0 Z \\ \hline \chi_4^0 & \stackrel{23\% }{\rightarrow} & \chi_1^+ W^- \\ & \stackrel{23\% }{\rightarrow} & \chi_1^- W^+ \\ & \stackrel{6\% }{\rightarrow} & \chi_1^0 h \\ & \stackrel{6\% }{\rightarrow} & \chi_1^0 Z \\ \hline \tilde{\tau}_1^+ & \stackrel{100\% }{\rightarrow} & \chi_1^0 \tau^+ \\ \hline \end{array} \end{array} \nonumber \end{eqnarray} \caption{Dominant decay channels and BRs of the final state (s)particles in $g g \rightarrow \tilde{q}_\chi \tilde{q}_{\chi '}^* X $, $X=h,H,A,H^\pm$, $q=t,b$ and $\chi,\chi'=1,2$, for $M_0=M_{1/2}=130$ GeV, $A_0=0$, $\tan\beta=40$ and ${\mathrm{sign}}(\mu)<0$ \cite{isajet}.} \label{table6} \end{table} \end{center} {\bf \ref{signatures}.2 Large $\tan\beta$ regime} \vskip0.5cm\noindent In the large $\tan\beta$ regime there is a variety of cross sections which can be significant: $\tilde{t}_1 \tilde{t}_1^* X$, $\tilde{t}_1 \tilde{t}_2^* X$, $\tilde{b}_1 \tilde{b}_1^* X$, $\tilde{b}_2 \tilde{b}_2^* X$, $\tilde{t}_1 \tilde{t}_2^* A$, $\tilde{b}_1 \tilde{b}_2^* A$, $\tilde{t}_1 \tilde{b}_1^* H^+$ and $\tilde{t}_1 \tilde{b}_2^* H^+$, where $X=h,H$. For reasons of space, however, we only focus our attention to one signature for the Higgs states not yet considered, the one arising from the dominant production channel in all cases, with the only exception of the charged Higgs bosons. In fact, we will neglect analysing here their decay patterns, as we have already mentioned the poor effectiveness of the charged Higgs production modes both as discovery channels and probes of the underlying M-SUGRA model. In the case of heavy scalar Higgs bosons, we consider the final state $\tilde{b}_1 \tilde{b}_1^* H$. This yields a cross section of 14 fb for $\tan\beta=40$. A possible decay chain is the one below. $$ \arraycolsep=0pt \begin{array}{lllllll}\label{Hchain} {\tilde{b}_1} && &~~~~{\tilde{b}}_1^* &&~~~~H~~~~~ & \\ \downarrow &&& ~~~\downarrow &&~~~~\downarrow~~~~~ & \\ \chi_1^0 \,+\,b\; & &&~~~~\chi_1^0 \,+\,\bar b &~&~~~~~b \,+\,\bar b & \end{array} $$ That is, a rather simple final state made up by four $b$-quarks and missing energy. The BR of this sequence is about 9\% (see Tab.~\ref{table6}). Therefore, at high luminosity, one obtains 129 events per year, before heavy flavour identification. The main background is certainly ordinary QCD production of four jets. However, the requirement of tagging four $b$-jets would reduce the latter considerably. Furthermore, if the mass of the heavy scalar Higgs boson is known, then one could impose that two $b$-quarks reproduce $m_H\approx 121$ GeV within a few GeV (say, 5) in invariant mass. Finally, given the enormous mass difference $m_{\tilde b_1}-m_b\approx 250$ GeV, compared to the rest mass of the LSP, $m_{\tilde \chi_1^0}\approx51$ GeV, one should expect, on the one hand, a large amount of missing energy, and, on the other hand, all four $b$-jets to be quite hard, both aspects further helping to reduce the QCD noise. In the end, some 32 events could well be detected annually, having already accounted for the overall $b$-tagging efficiency $\varepsilon_b^4=0.25$. For the case of the pseudoscalar Higgs particle, we choose the final state ${\tilde t}_1 {\tilde t}_2^* A$, which has a cross section of about 79 fb at $\tan\beta=40$. A possible signature could be\footnote{Additional examples can be found in Ref.~\cite{short}.}: $$ \arraycolsep=0pt \begin{array}{lllllll}\label{Achain} {\tilde{t}_1} && &~~~~{\tilde{t}}_2^* &&~~~A~~~~~ & \\ \downarrow &&& ~~~\downarrow &&~~~\downarrow~~~~~ & \\ \chi_1^+ \,+\,b\; & &&~~~~\chi_1^- \,+\,\bar b &~&~~~~b \,+\,\bar b & \\ \downarrow &&& ~~~\downarrow && \; & \\ {\tilde \tau}_1 \,+\,\nu& && ~~~~{\tilde \tau}_1^* \,+\,\bar\nu &&&\\ \downarrow &&& ~~~\downarrow && \; & \\ \tau^+\,+\,\chi_1^0& && ~~~~\tau^-\,+\,\chi^0_1 &&& \end{array} $$ Here, the final state is made up by four $b$-quarks and two $\tau$'s, plus missing energy as usual. The decay fraction is 23\% (again, see Tab.~\ref{table6}). That is, 1814 events per year before tagging $b$'s and $\tau$'s. The most dangerous backgrounds are probably $Z+4$~jet production and $t\bar t b\bar b$ events. The first can be rejected by asking, e.g., $M_{\tau^+\tau^-}\ne M_Z$, if $\tau$'s are reconstructed. In addition, both background processes have (at least) two $b$-quarks quite soft. As for the signal, given that the lightest chargino mass is much smaller than the stop ones (in fact, $m_{{\tilde\chi}^\pm_1}\approx93$ GeV whereas $m_{{\tilde t}_1}\approx248$ GeV and $m_{{\tilde t}_2}\approx388$ GeV), all $b$'s are naturally energetic and two of them also peak at $m_A\approx120$ GeV. Thus, to require all $M_{bb}$ invariant masses sufficiently large with one close to the $A$ mass should help in enhancing considerably the signal-to-background rates. Requiring large missing energy would help further, especially against $t\bar t b\bar b$ events. More difficult is to discern differences in the $\tau$ behaviours (though, notice that $m_{{\tilde \tau}_1}\approx76$ GeV $\gg m_\tau$). For $\varepsilon_b^4=0.25$, and assuming leptonic decays of both $\tau^+$ and $\tau^-$ into electrons and/or muons, one finally gets something of the order of 110 signal events per year. \subsection{Heavy mass spectrum} \label{heavyspectrum} An attempt to summarise our findings is made in Fig.~\ref{fig:light}, where the most relevant cross sections (see upper frame) for the light mass regime (see lower frame) are plotted for a choice of $M_{0}$, $M_{1/2}$, $A_0$, $\tan\beta$ and $\mathrm{sign}(\mu)$ which reflects their hierarchal order seen over most of the M-SUGRA parameter space discussed so far. However, to assume that only light squark and Higgs masses can induce sizable cross sections in events of the type (\ref{proc}) would be wrong. This is a sufficient condition for many channels, but not a necessary one. For example, even for very large universal masses, one could find a value of the soft trilinear coupling small enough to overcome the loss of signal due to propagator and phase space effects. In fact, as repeatedly shown in the previous Section, most of the cross sections considered here grow quickly as $A_0$ becomes negative. Fig.~\ref{fig:heavy} makes eloquently this point (top insert), for the choice $A_0=-900$ GeV, well consistent with the bounds imposed by the charge and colour breaking minima. There, we have adopted a very large value for $M_0$, i.e., 500 GeV, and varied $M_{1/2}$ between 100 and 500 GeV. The choice of a large $\tan\beta$ value, i.e., 35, is necessary to obtain detectable rates, except in those cases involving ${\tilde t}_1$ and $h$ in the same event. In contrast, that of a negative $\mathrm{sign}(\mu)$ never is. The squark and Higgs masses produced by the above combinations of M-SUGRA parameters can be found in the bottom frame of Fig.~\ref{fig:heavy}. There are only a few production channels which survive the strong phase space suppression arising in the heavy mass regime and yield cross sections above 1 fb. Among these, other than those already encountered ${\tilde t}_1{\tilde t}_1^* h$, ${\tilde t}_1{\tilde t}_2^* h$, ${\tilde b}_1{\tilde b}_1^* h$, one notices the appearance of channels which had negligible rates in the low mass regime, notably ${\tilde b}_1{\tilde b}_2^* h$ (compare to Fig.~\ref{fig:lighthiggs}). In this specific instance, the effect is due to the enhancement induced by the onset of the $\tilde{b}_2 \rightarrow \tilde{b}_1 h$ decay mode in $gg\to \tilde{b}_2 \tilde{b}_2^*$ events. Even the other two combinations ${\tilde t}_1{\tilde t}_1^* H$ and ${\tilde t}_1{\tilde b}_2^* H^-$, which had a rather low profile in the light mass regime, now compete more closely with the dominant modes. Finally, notice that pseudoscalar Higgs boson production is no longer significant in this mass regime, in line with the results presented in Ref.~\cite{short}. In practise, if $M_0=500$ GeV, events involving light CP-even Higgs bosons could be detected up to $M_{1/2}=400$ GeV or so, in either mode ${\tilde t}_1{\tilde t}_1^* h$ or ${\tilde t}_1{\tilde t}_2^* h$. Final states of the type ${\tilde b}_1{\tilde b}_2^* h$ have surprisingly large rates if $M_{1/2}$ is below 220 GeV. The maximum reach in $M_{1/2}$ via ${\tilde b}_1{\tilde b}_1^* h$, ${\tilde t}_1{\tilde b}_2^* H^-$ and ${\tilde t}_1{\tilde t}_1^* H$ is instead 220, 180 and 140 GeV, respectively. This is just one example where a new phenomenology of squark-squark-Higgs events arises for rather heavy $M_0$ and $M_{1/2}$ masses. We have found several such combinations, and linked them to the fact that $A_0$ ought to be significantly large and negative, $\tan\beta$ close to $m_t/m_b$, but with small dependence on $\mathrm{sign}(\mu)$. \section{Conclusions} \label{sec:conclusions} In summary, we have studied neutral and charged Higgs boson production in association with all possible combinations of stop and sbottom squarks at the LHC, in the context of the SUGRA inspired MSSM. Our interest in such reactions was driven not only by the fact that they can act as new sources of Higgs particles but also because they carry a strong dependence on the five inputs of the SUSY model, so that they can possibly be used to constrain the latter. In a sense, this note (along with Ref.~\cite{short}) completes previous analyses on the subject \cite{djouadi,djouadi2,fawzi}, where the emphasis was mainly put on the usefulness of the above kind of reactions as Higgs production modes and the attention consequently restricted to the case of light squark and Higgs masses. We have found that the cross sections of many of these processes should be detectable at high collider luminosity for not too small values of $\tan\beta$. Indeed, their production rates are strongly sensitive to the ratio of the VEVs of the Higgs fields, this possibly allowing one to put potent constraints on such a crucial parameter of the MSSM Higgs sector. Furthermore, also the trilinear coupling $A_0$ intervenes in these events, in such a way that visible rates would mainly be possible if this other fundamental M-SUGRA input is negative. (Indeed, to know the actual value of $\tan\beta$ from other sources would further help to assess the magnitude of $A_0$.) As for the sign of the Higgsino mass term, $\mathrm{sign}(\mu)$, it affects the phenomenology of such events in one or two channels only, so that it can easily evades the imposition of experimental bounds. Finally, concerning the remaining two parameters (apart from mixing effects) of the M-SUGRA scenario, one must say that $M_0$ need not be small (it could be as large as 500 GeV) and that $M_{1/2}$ is enough to be below 220 GeV in order to guarantee sizable cross sections in many cases. In a few representative examples, we have further investigated the decay phenomenology of these reactions, by discussing some possible signatures, their rates (of the order of tens to hundreds of events per year at high luminosity) and peculiar kinematics, as opposed to the yield of ordinary, non-SUSY backgrounds. In conclusion, we believe these processes to be potentially very helpful in putting drastic limits on several M-SUGRA parameters and we thus recommend that their phenomenology is further investigated in the context of dedicated experimental simulations, which were clearly beyond the scope of this note. In this spirit, we have derived compact analytical formulae of the relevant production MEs, that can easily be incorporated in existing MC programs. \section*{Acknowledgements} S.M. acknowledges the financial support from the UK PPARC and useful conversations with Michael Kr\"amer. A.D is supported from the Marie Curie Research Training Grant ERB-FMBI-CT98-3438 and thanks Mike Seymour for useful discussions. We both thank Herbi Dreiner for precious comments and The Old School in Oxford for kind hospitality.
\section{Basics} Consider $N({\bf p}) N(- {\bf p}) \rightarrow N ({\bf p}') N (-{\bf p}')$ scattering in the $^1S_0$ channel. Since the spins of the two nucleons are combined anti-symmetrically Fermi statistics implies that this channel is $I = 1$ (similarly the $^3S_1$ and $^3D_1$ channels are $I = 0$). The energy $E = p^2/M = p^{\prime 2}/M$ where $p^{(\prime)} = |{\bf p}^{(\prime)}|$ and the scattering matrix, $S$, is related to the scattering amplitude ${\cal A}$ by $S = 1 + i Mp {\cal A} /2\pi$. Since $S = e^{2i\delta}$, where $\delta$ is the phase shift, \begin{equation}\label{1} {\cal A}^{(^1S_0)} = \frac{4\pi}{M} \frac{1}{p\cot \delta^{(^1S_0)} - ip}, \end{equation} where $M$ is the nucleon mass. For $p< m_\pi/2$ the quantity $p\cot\delta$ can be expanded in a power series in $p^2$ \begin{equation}\label{2} p\cot\delta^{(^1S_0)} = - \frac{1}{a^{(^1S_0)}} + \frac{1}{2} r_0^{(^1S_0)} p^2 + \ldots , \end{equation} when $a$ is called the scattering length and $r_0$ is called the effective range. The scattering length in the $^1S_0$ channel is very large,~\cite{burcham} $a^{(^1S_0)} \simeq - 23.7$ fm or $1/a^{(^1S_0)} \simeq - 8.3$ MeV. On the other hand the nuclear potential is characterized by a momentum scale $\Lambda \sim 200$ MeV. The smallness of $|1/a^{(^1S_0)}|$ compared with this scale is the result of an accidental cancellation which causes a state in the spectrum to be very near zero binding energy. ($a \rightarrow - \infty$ as a scattering state approaches zero energy and $a \rightarrow \infty$ as a bound state approaches zero binding energy.) Neglecting the small $^3S_1- ^3D_1$ mixing, formulas analogous to eqs.~(\ref{1}) and~(\ref{2}) hold in the $^3S_1$ channel. The scattering length is also large in that case,~\cite{burcham} $a^{(^3S_1)} \simeq 5.4$ fm or $1/a^{(^3S_1)} \simeq 36$ MeV. The bound state in this channel that is near zero binding energy is the deuteron. \section{Expansions of ${\cal A}$} The simplest expansion of ${\cal A}$ is a momentum expansion. This is analogous to what is done in standard applications of effective field theory, e.g., chiral perturbation theory for $\pi\pi$ scattering. For $NN$ scattering in the $s =~ ^1\! S_0$ or $^3S_1$ channels, \[ {\cal A}^{(s)} = \frac{4\pi}{M} \frac{1}{\left[- 1/a^{(s)} + \frac{1}{2} r^{(s)}_0p^2 + \ldots - ip\right]}\] \begin{equation}\label{3} = - \frac{4\pi}{M} a^{(s)} \left\{1 - i a^{(s)} p + \left(\frac{a^{(s)} r_0^{(s)}}{2} - a^{(s)^{2}}\right) p^2 + \ldots\right\}. \end{equation} If $a^{(s)}$ was its natural size (i.e., $a^{(s)} \sim 1/\Lambda$) this would be the appropriate expansion to perform. However, in nature the $S$-wave $NN$ scattering lengths are very large and the expansion above is only valid in the small region of momentum $p\lsim |1/a^{(s)}| \ll 1/\Lambda.$ Since the underlying physics is set by $m_\pi$ and $\Lambda_{QCD}$ there should be an expansion in $p/\Lambda$ that is valid even when $p \gg |1/a^{(s)}|$. It is not difficult to deduce what this expansion is. In Eq.~(\ref{3}) keep $-1/a^{(s)} - ip$ in the denominator and expand in the remaining terms. This yields \begin{equation}\label{4} {\cal A}^{(s)} = - \frac{4\pi}{M} \frac{1}{(1/a^{(s)} + ip)} \left[1 + \frac{r_0^{(s)} p^2/2}{(1/a^{(s)} + ip)} +\ldots\right]. \end{equation} Now ${\cal A}^{(s)} = \sum_{n = - 1}^{\infty} {\cal A}_n^{(s)}$, where ${\cal A}_n^{(s)} \sim {\cal O} (p^n)$. This is the appropriate expansion in the case where the scattering lengths are large. It has the unusual property that the leading term is order $p^{-1}$. \section{Effective Field Theory Without Pions} The effective field theory with the pions integrated out contains only nucleon fields, $N = \left({p\atop n}\right)$, and we expect that the lowest dimension operators will be the most important ones. The Lagrange density is written as, ${\cal L} = {\cal L}_1 + {\cal L}_2 + \ldots,$ where ${\cal L}_n$ contains $n$-body operators. The one and two body terms are: \begin{equation}\label{5} {\cal L}_1 = N^\dagger \left[i \partial_t + \frac{\vec \nabla^2}{2M}\right] N + \ldots, \end{equation} \begin{equation}\label{6} {\cal L}_2 = - \sum_s C_0^{(s)} (N^T P_i^{(s)} N)^\dagger (N^T P_i^{(s)} N) + \ldots . \end{equation} Here $s = ~^1S_0$ or $^3S_1$, the ellipses denote higher dimension operators and $P_i^{(s)}$ are the spin-isospin projectors \begin{equation}\label{7} P_i^{(^1S_0)} = \left(\frac{\sigma_2 \tau_2 \tau_i}{\sqrt{8}}\right), P_i^{(^3 S_1)} = \left(\frac{\sigma_2 \sigma_i \tau_2}{\sqrt{8}}\right), \end{equation} where the Pauli matrices $\sigma_i$ act in spin space and the Pauli matrices $\tau_i$ act in isospin space. \begin{figure}[t!] \centerline{\psfig{figure=diag0.eps,width=17.0cm}} \caption[1]{The leading contribution to $NN$ scattering.} \label{fig_C00} \end{figure} Neglecting higher dimension operators the scattering amplitudes in the $^1S_0$ and $^3 S_1$ channels come from the sum of bubble-type Feynman diagrams shown in Fig. 1. Each bubble is linearly divergent in the ultraviolet so the coefficients $C_0^{(s)}$ depend on the regulator and subtraction scheme adopted. We use dimensional regularization and start with minimal subtraction (we will switch to a different subtraction scheme momentarily). Since the divergences are linear the Feynman diagrams have poles at $D = 3$ but not at $D = 4$. In $MS$ (minimal subtraction) the coefficients of the operators explicitly displayed in Eq.~(\ref{6}), are subtraction point independent and we denote them by $\bar C_0^{(s)}$. In this scheme the sum of bubble-type Feynman diagrams gives \begin{equation}\label{8} {\cal A}^{(s)} = - \frac{ \bar C_0^{(s)}}{1 + iMp \bar C_0^{(s)}/4\pi}. \end{equation} Comparing Eq.~(\ref{8}) with eqs.~(\ref{1}) and~(\ref{2}) it is evident that this corresponds to keeping only the scattering length term in the expansion of $p\cot \delta^{(s)}$, (i.e., the first term of Eq.~(\ref{4})) and that \begin{equation}\label{9} \bar C_0^{(s)} = \frac{4\pi a^{(s)}}{M}. \end{equation} So in this subtraction scheme the coefficients $\bar C_0^{(s)}$ are very large and also very different in the two channels. However as $a^{(s)} \rightarrow \infty, {\cal A}^{(s)} \rightarrow 4\pi i/Mp$ which is the same in both channels. This form for the scattering amplitudes is consistent with Wigner spin-isospin $SU(4)$ symmetry, and also with scale invariance. In MS when $p> 1/a^{(s)}$ the terms in the perturbative series for the scattering amplitude get larger and larger. We would like to use a subtraction scheme where the various Feynman diagrams in Fig. 1 are the same size as their sum and where the symmetries that arise as $a^{(s)} \rightarrow \infty$ are manifest at the level of the Lagrangian. Examples of such subtraction schemes are~\cite{kaplan1} PDS where poles at $D=3$ are also subtracted and the $OS$ momentum space subtraction scheme.~\cite{weinberg1,mehen1} In these schemes the coefficients are subtraction point dependent, $C_0^{(s)} = C_0^{(s)} (\mu)$, and the sum of bubble diagrams gives \begin{equation}\label{10} {\cal A}^{(s)} = - \frac{C_0^{(s)} (\mu)}{1 + M (\mu + ip) C_0^{(s)} (\mu)/4\pi}. \end{equation} This still corresponds to keeping just the scattering length, and is the leading term in Eq.~(\ref{4}). But now \begin{equation}\label{11} C_0^{(s)} (\mu) = - \frac{4\pi}{M} \frac{1}{\mu - 1/a^{(s)}}, \end{equation} which as $a^{(s)} \rightarrow \infty$ becomes $C_0^{(s)} (\mu) = - 4\pi/M\mu$. In this limit, the coefficients are the same in both channels and with $\mu \sim p$ each term in sum of bubble type Feynman diagrams in Fig. 1 is the same size as the sum itself. The operators with coefficients $C_0^{(s)}$ are nonrenormalizable dimension six operators. Naively they are irrelevant operators and at low momentum can be treated in perturbation theory. However as $a^{(s)} \rightarrow \infty$ the coefficients $C_0^{(s)} (\mu)$ flow to a nontrivial fixed point~\cite{kaplan1,weinberg1} where $\mu d [\mu C_0^{(s)} (\mu)]/d\mu = 0$. For large $a^{(s)}$ the power counting is controlled by this fixed point and the leading contribution to the $NN$ scattering amplitude comes from treating $C_0^{(s)}$ nonperturbatively. It is straightforward to show that in PDS or OS the coefficients of $S$-wave operators with $2n$ spatial derivatives scale as,~\cite{kaplan1} \begin{equation}\label{12} C_{2n}^{(s)} (\mu) \sim \frac{4\pi}{M\Lambda^n\mu^{n+1}}, \end{equation} for $\mu \gg |1/a^{(s)}|$. With $\mu \sim p, C_{2n}^{(s)} (\mu) p^{2n} \sim p^{n-1}$ and the two body operators with derivatives can be treated perturbatively. In a non-relativistic theory a loop integration $\int d^4 q = \int dq^0 d^3 q \sim {\cal O} (p^5)$ (since the $dq^0$ integration is of order $p^2$ and the $d^3q$ is order $p^3$) and the nucleon propagator $i/(p^0 - p^2/2M + i\epsilon) \sim {\cal O} (p^{-2})$. Consequently each loop gives a factor $p$ plus whatever factors of $p$ are associated with the vertices. The power counting~\cite{kaplan1,kolck} is now evident. The leading order $(LO)$ contribution ${\cal A}_{-1}^{(s)}$ comes from $C_0^{(s)}$ treated nonperturbatively, the next to leading order $(NLO)$ contribution ${\cal A}_0^{(s)}$ comes from $C_0^{(s)}$ treated nonperturbatively and $C_2^{(s)}$ inserted once, the next-to-next to leading order $(N^2 LO)$ contribution comes from $C_0^{(s)}$ treated nonperturbatively, $C_2^{(s)}$ inserted twice or $C_4^{(s)}$ inserted once, etc. With the pions integrated out the effective field theory expansion applied to $NN$ scattering reproduces Eq.~(\ref{4}) and has no more content than the momentum expansion of $p\cot\delta^{(s)}$. However, even with the pions integrated out one can couple photons or $W$ and $Z$ gauge bosons to the nucleons. The relative importance of operators containing these fields depends on their renormalization group scaling near the fixed point. In the two nucleon sector predictions based on the effective field theory without pions are similar to those made by effective range theory.~\cite{bethe} However the effective field theory approach has a number of advantages. Predictions based on effective range theory are only valid to a given order in the $p/\Lambda$ expansion. In the effective field theory new two-body operators containing the gauge fields arise which spoil the predictions of effective range theory. For the thermal neutron capture cross section, $\sigma(n + p \rightarrow d + \gamma)$, this occurs at $NLO$ while for the deuteron matter (charge) radius $<r_m>$ this doesn't occur until $N^3LO$. This explains why the effective range theory prediction for $\sigma(n + p \rightarrow d + \gamma)$ is off by 10\% while the effective range theory prediction for $<r_m>$ is accurate to better than a percent.\cite{chen} For these static deuteron properties the relevant momentum in the $p/\Lambda$ expansion is set by the deuteron binding energy, i.e., $p \sim 40$ MeV. Another useful aspect of the effective field theory formalism is that it is straightforward to include relativistic corrections. As $a^{(s)} \rightarrow \infty, {\cal L}_2 \rightarrow - (2\pi/M\mu) (N^\dagger N)^2 + \ldots$, where the ellipses denote two body operators with derivatives. In this limit the leading one and two body terms are invariant under the following symmetries:~\cite{mehen2} \begin{description} \item{(i.)} Wigner Symmetry~\cite{wigner}\\ Under infinitesimal Wigner symmetry $SU(4)$ transformations \begin{equation}\label{13} \delta N = i \alpha_{\mu\nu} \sigma^\mu \tau^\nu N, \end{equation} where $\sigma^\mu = (1,\mbox{\boldmath $\sigma$})$ and $\tau^\mu = (1, \mbox{\boldmath $\tau$})$ with $\mu = 0,1,2,3$ and repeated indices summed. The symmetry group corresponding to Eq.~(\ref{13}) is actually $SU(4)\times U(1)$, with $\alpha_{00}$ the group parameter for the additional baryon-number $U(1)$. Associated with this symmetry are the conserved charges, \begin{equation}\label{14} Q^{\mu\nu} = \int d^3 x N^\dagger \sigma^\mu\tau^\nu N. \end{equation} The two body terms with derivatives are not invariant under Wigner symmetry even if $a^{(s)} \rightarrow \infty$. Hence in the two body sector the violations of Wigner symmetry go as, $(1/[a^{(^1S_0)}p] - 1/[a^{(^3S_1)} p])$ and $p/\Lambda$. Wigner symmetry will not be a good approximation if the momentum $p$ is too low or if it is too large. Wigner symmetry is relevant for nuclei with many nucleons.~\cite{group} It is not difficult to see that the higher body terms with no derivatives are automatically invariant under Wigner symmetry. Since these contact terms are antisymmetric in the nucleon fields $N$ and in the hermitian conjugates $N^\dagger$, contact terms without derivatives cannot occur for five body operators and higher. The nucleons $N$ are in the ${\bf 4}$ of $SU(4)$ and the $N^\dagger$'s are in the $\bar{\bf 4}$. Four nucleons combined anti-symmetrically are an $SU(4)$ singlet and so the four-body terms are invariant under $SU(4)$. The three body terms transform as $\bar{\bf 4} \otimes {\bf 4} = {\bf 1} \otimes {\bf 15}$. However the operators in the ${\bf 15}$ are not invariant under the total spin or isospin $SU(2)$ subgroups of $SU(4)$. Hence the allowed three body terms are also invariant under $SU(4)$ Wigner symmetry. A complete extension of the general fixed point power counting to the higher body terms has not been made. However there has been considerable recent progress.~\cite{kolck} This work indicates that the 3-body term with no derivatives is leading order (i.e., as important as effects coming from $C_0^{(s)}$). \item{(ii.)} Scale Invariance\\ The leading one and two body terms are invariant under the scale transformation $N(t,{\bf x}) \rightarrow N' (t,{\bf x})$ and $\mu \rightarrow \mu'$ where \begin{equation}\label{15} N' (t,{\bf x}) = \lambda^{-3/2} N (t/\lambda^2, {\bf x}/\lambda), \end{equation} \begin{equation}\label{16} \mu' = \mu/\lambda. \end{equation} Note that Eq.~(\ref{15}) corresponds to $N'(t', {\bf x}') = \lambda^{-3/2} N(t, {\bf x})$ with ${\bf x}' = \lambda {\bf x}$ and $t' = \lambda^2 t$. The different scaling of space and time coordinates is dictated by invariance of the leading one-body terms in the Lagrange density. \end{description} \section{Including Pions} With pions included the power counting is taken to be in powers of $Q/\Lambda_{NN}$ where $p\sim m_\pi \sim Q$. A subscript $NN$ has been put on $\Lambda$ as a reminder that the expansion should work better if the pions are included as explicit fields, i.e., we expect that $\Lambda_{NN} > \Lambda$. Potential pion exchange arises from the term \begin{equation}\label{17} {\cal L}_{{\rm int}} = - \frac{g_A}{\sqrt{2} f_\pi} \nabla^i \pi^j N^\dagger \sigma^i \tau^j N, \end{equation} where $g_A \simeq 1.25$ is the axial coupling and $f_\pi \simeq 131$ MeV is the pion decay constant. Exchange of a potential pion between nucleons is order $Q^0$ (the two factors of $Q$ from the vertices cancel the $1/Q^2$ from the pion propagator). This is the same size as the two body contact terms with two derivatives and consequently pion exchange can be treated perturbatively. Including pion exchanges without the two derivative two body contact terms is not a systematic improvement and is no better (from a power counting perspective) than just including the effects of $C_0^{(s)}$. Note that this power counting is very different from the one originally proposed by Weinberg~\cite{weinberg1,weinberg2} where the leading contribution came from treating both potential pion exchange and $C_0^{(s)}$ nonperturbatively. The effects of two body terms with derivatives and insertions of the light quark mass matrix were considered subdominant. Weinberg's power counting treats the nucleon mass $M$ as large and $MQ \sim {\cal O}(1)$. It assumes that factors of $M$ only arise from the loop integrations. In a toy model where perturbative matching between the full relativistic theory and the nonrelativistic effective theory can be explicitly performed Luke and Manohar~ {\cite{luke}} found that the two body local operators in the effective nonrelativistic theory have coefficients that contain factors of M. This is the origin of the problem with Weinberg's power counting. \begin{figure}[t!] \centerline{\psfig{figure=diag1.eps,height=3.0cm,rwidth=16cm}} \caption[1]{Contribution to $^3S_1$ scattering from three potential pion exchange. The dashed lines denote potential pion exchange.} \label{fig_ct0pi} \end{figure} In the $^3S_1$ channel the Feynman diagram shown in Fig. 2 with three potential pion exchanges is logarithmically divergent. Neglecting the pion mass it gives a contribution to ${\cal A}^{(^3S_1)}$ of order \begin{equation}\label{18} \left(\frac{4\pi}{M}\right) \left(\frac{Mg_A^2}{8\pi f_\pi^2}\right)^3 p^2 [\ln \mu^2 + K], \end{equation} where $K$ is a constant. The $\mu$-dependence above is cancelled by the $\mu$-dependence of $C_2^{(^3S_1)}$. There is no point to including this Feynman diagram without including the effects of the two body $^3S_1$ operator with 2-derivatives. Eq.~(\ref{18}) is an $N^3LO$ contribution. With the pions included a single insertion of $C_2^{(s)}$ is not just $NLO$ it contributes at higher levels at the $Q$ expansion as well. For that reason $C_2^{(s)}$ and the other contact term coefficients are sometimes written as a sum $C_2^{(s)} = \sum_{a = 1}^\infty C_{2,a}^{(s)}$ where $C_{2,1}^{(s)}$ gives the $NLO$ contribution, etc. When this is done predictions for physical quantities are exactly $\mu$ independent, at each order in the $Q$ expansion. If $C_2^{(s)}$ is not expanded in this way then predictions at a given order on the $Q$ expansion have some subtraction point dependence, which is higher order in the $Q$ expansion. \begin{figure}[t!] \centerline{\psfig{figure=diag2.eps,height=3.0cm,rwidth=16cm}} \caption[1]{Contribution to $NN$ scattering that renormalizes $D_2^{(s)}$.} \label{fig_ct1pi} \end{figure} There are two body $S$-wave contact terms with no derivatives but with an insertion of the light quark mass matrix, \begin{equation}\label{19} m_q = \left(\begin{array}{cc} m_u & 0\\ 0 & m_d \end{array} \right). \end{equation} Since $m_\pi^2 \propto (m_u + m_d)$ an insertion of $m_q$ counts as two powers of $Q$ and the coefficients of these operators $D_2^{(s)}$ scale with $\mu$ in the same way as the coefficients $C_2^{(s)}$. At $NLO$ they must also be included. The Feynman diagram in Fig. 3 is logaritmically divergent and it gives a contribution to the $^1S_0$ and $^3S_1$ scattering amplitudes of order \begin{equation}\label{20} \left(\frac{4\pi}{M}\right) \left(\frac{g_A^2 M}{8\pi f_\pi^2}\right) \left(\frac{C_0^{(s)} M}{4\pi}\right)^2 m_\pi^2 [\ln \mu^2 + K], \end{equation} where $K$ is a constant. The $\mu$ dependence here is cancelled by that of the coefficients $D_2^{(s)}$. Including one pion exchange without the effects of the two body terms with one insertion of the quark mass matrix does not systematically improve the theoretical prediction for the $NN$ scattering amplitude. If a momentum cutoff regulator is used instead of dimensional regularization then including pion exchange without the two body contact operators that have an insertion of the quark mass matrix results in a cutoff dependent amplitude ${\cal A}^{(s)}$. It is possible in the $^1S_0$ channel to sum to all orders potential pion exchange and when this is done the cutoff dependence does not become subdominant~\cite{kaplan2} (compared with the finite cut off independent parts of pion exchange). The effects of local four nucleon (i.e., two body) operators with an insertion of the quark mass matrix cannot be viewed as less important than the effects of pion exchange. The conventional explanation for the discrepancy between the prediction of effective range theory for the thermal neutron capture cross section $\sigma (n + p \rightarrow d + \gamma$) and its experimental value is meson exchange currents,~\cite{riska} which roughly speaking are the contribution of Feynman diagrams where the photon couples to a potential pion. In the effective field theory approach with the pions included this discrepancy is made up~\cite{savage} (at least partly) by the $NLO$ contribution which involves both meson exchange current Feynman diagrams and the contribution of a local two body operator involving the magnetic field. \section{An Application of Wigner Symmetry} Potential pions have $k^0 \sim {\bf k}^2/M$ while radiation pions have $k^0 \sim \sqrt{{\bf k}^2 + m_\pi^2}$. The coupling of the radiation pions to the nucleons is done by performing a multipole expansion on Eq.~(\ref{17}). At leading order this amounts to evaluating the pion field in Eq.~(\ref{17}) at the space time point $(t,{\bf x}) = (t, {\bf 0})$. Hence, for radiation pions the term in the action corresponding to Eq.~(\ref{17}) is \begin{equation}\label{21} S_{int} = - \frac{g_A}{\sqrt{2} f} \int dt (\nabla^i \pi^j)|_{{\bf x} = 0} Q^{ij}, \end{equation} where $Q^{ij}$ are the charges of Wigner symmetry in Eq.~(\ref{14}). In the limit $a^{(s)} \rightarrow \infty$ these charges are conserved and the $Q^{ij}$'s are time independent. Hence, as $a^{(s)} \rightarrow \infty$ only the $k^0 = 0$ mode of the pion couples in Eq.~(\ref{21}). This is incompatible with the radiation pion condition, $k^0 \sim \sqrt{{\bf k}^2 + m_\pi^2}$. Hence the leading contribution from radiation pions is suppressed by $1/a^{(^1S_0)} - 1/a^{(^3S_1)}$. In a recent paper~\cite{mehen3} Mehen and Stewart found this suppression by an explicit calculation of the leading radiation pion contribution to the $NN$ scattering amplitudes, ${\cal A}^{(s)}$. It involved a cancellation between different Feynman diagrams. \section{Outlook} Effective field theory methods are a viable model independent approach to the physics of the two nucleon sector. The power counting is slightly unusual due to the large $S$-wave $NN$ scattering lengths. This approach is useful up to a center of mass momentum around $200~{\rm MeV}$, however, the expansion parameter at such a momentum is probably not much smaller than ${1 \over 2}$. It seems likely that for many quantities calculations at $N^2LO$ will reach the same precision as conventional potential model approaches, however, with such a large expansion parameter there are likely to be some failures. Extension of the effective field theory approach to the three nucleon sector is underway. Several theoretical issues remain to be resolved before there is a complete power counting, but recent progress in this area is very encouraging. The holy grail of this field is the application of these effective field theory methods to nuclear matter. We are still a long way from having the theoretical tools to tackle this problem and even with these tools the Fermi momentum associated with nuclear density may be too large for a $Q$ expansion to be useful. However, given the importance of understanding the properties of nuclear matter continuing to develop the effective field theory approach is very worthwhile. This work was supported in part by the Department of Energy under grant number DE-FG03-92-ER 40701. \section*{References}
\section*{Introduction and definition of the model} The Edwards-Anderson Model for spin-glasses is one of the most intensively studied models in the domain of disordered systems. It combines a very simple formulation and rich behavior. However in its standard form it does not possess a finite temperature transition in dimension two \cite{bhatt1988}, so the study of finite temperature spin glass phase transitions is restricted to three dimensions or above. This is unfortunate as dimension two has many advantages, including the facility for direct visualization. The Random Field Ising Model (RFIM) proposed by Imry and Ma \cite{imry1975} is also a very important model for disordered systems. Unfortunately this model does not present a phase transition in two dimensions \cite{imbrie1984}, implying that once more we have to go to higher dimensions to study a phase transition. These results seem to imply that no intrinsically two dimensional disordered system present a finite transition temperature. This statement has been questioned in recent years and some disordered systems have been proposed for a finite temperature phase transition in two dimensions \cite{shirakura1997,pasquini1997}. We introduced \cite{lemke1996b} a model for Ising spins on a square lattice where second neighbors are coupled ferromagnetically with an interaction strength $J$, and where there is a random near neighbor coupling of strength $\pm \lambda J$. The model is described by the following Hamiltonian: \begin{equation} \label{def.eq} {\cal H} =\sum_{\langle i,j \rangle} - S_iJ S_j + \sum_{\{ i,j\} } S_iJ_{ij}S_j \end{equation} where $\langle i,j \rangle $ means a sum over second neighbors, $\{ i,j\}$ over first neighbors and $J_{ij}=\pm \lambda J$. We have decided to call this model the Randomly Coupled Ferromagnet (RCF). We presented simulations which indicated the presence of a phase transition at a finite temperature $T_c$ near $2$ (in units of $J$) for $\lambda < 1$. Similar behavior were also found on the analogous XY model \cite{jain1998}. The conclusions we drew for the Ising version of this model were contested in a paper by Parisi, Ruiz-Lorenzo, and Stariolo \cite{parisi1998}. These authors carried out simulations to larger sizes, up to $L=48$. They interpreted their data in terms of size dependent crossovers at low temperatures, successively between staggered ferromagnet, spin-glass like, and paramagnetic phases as size $L$ increases. In \cite{parisi1998} a picture for the low temperature phase was proposed implying that at large sizes the system is equivalent to a standard two dimension spin-glass, having no true finite temperature phase transition. Here we give arguments leading to a different picture for the low temperature phase. Simulations are presented which clearly indicate a phase transition at a finite temperature for a wide range of $\lambda$. \section*{General Discussion of the model} When $\lambda =0$ we simply have two independent sub-lattices $A$ and $B$ that order ferromagnetically and independently at the Onsager value of the Curie temperature, $T_c=2.26\ldots$. Below this temperature each sub-lattice has its overall magnetization either up or down; thus there are four different degenerate ground states. For finite values of $\lambda$, each sub-lattice will exert a random field on the other, so we expect that for a large enough lattice the long range ferromagnetic order in each sub-lattice will be destroyed, following the Imry-Ma argument for the two dimensional RFIM \cite{imry1975}. On each sub-lattice the system will be broken up into several different locally ferromagnetic domains whose size will be roughly given by \begin{equation} R_c\sim \exp\left[\frac{C}{\lambda^2} \right], \end{equation} where $C$ is a constant \cite{parisi1998,binder1983,seppala1998}. However here in contrast to the RFIM the ``effective random fields'' are not fixed once and for all but fluctuate in time as the spins in the other sub-lattice relax. The crucial point is: are the domains ``stable'' in time below some critical temperature? Alternatively, are they in perpetual motion at all finite temperatures, so that after a sufficient time all memory of an initial equilibrium spin configuration will be wiped out, even in the thermodynamic limit? In reference \cite{parisi1998} it is proposed that each of these domains can be regarded as a ``super-spin'' having finite random interactions with its neighbors. So the system will behave as ferromagnet for small scales and as a spin glass for larger scales. This picture implies that in the thermodynamical limit the system will only present a phase transition at $T=0$. We argue that this picture is incorrect, since in fact the ``super-spins'' do not behave as na\"{\i}vely expected. Let us first go to the RFIM limit. In figure \ref{fixed.fig} we present the results of a simulation at $\lambda = 0.5$ where the spins on sub-lattice $A$ are all frozen up, and those on sub-lattice $B$ can evolve following the Hamiltonian (\ref{def.eq}). Sub-lattice $A$ is reduced basically to one super-spin, and the figure also represents a snapshot of the $B$ sub-lattice configuration after a long anneal at temperature $T=1.5$, together with the time dependence of the auto correlation function after anneal, $q_B(t)$, at the same temperature for the $B$ spins. $q_B(t)$ is defined as: \begin{equation} \label{q.eq} q_B(t) =\lim_{t_w\to \infty}\sum_{i\in B} [ S_i(t_w)S_i(t_w+t) ] \end{equation} where $[ \ldots ]$ represent a configurational average. As we can see $q_B(t)$ quite rapidly reaches an asymptotic value, which is about 0.85 at this temperature. The system can be exactly described by the Hamiltonian: \begin{eqnarray} \label{rf.eq} {\cal H}(S_i=1 |\, i\in A) &=& \sum_{<i,j>} J S_i S_j + \sum_{i,j}S_iJ_{ij}S_j \\ &=& \sum_{<i,j>} J S_i S_j + \sum_{i}h_i S_i \end{eqnarray} where $h_i$ is a random field. This is precisely the RFIM hamiltonian. The Imbrie result \cite{imbrie1984} implies that for a infinite system $q_B(t)$ will tend to definite positive value for all temperatures. \ifthenelse{\equal{condmat}{condmat}}{ \begin{figure} \begin{center} \subfigure[]{\includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig1a.ps}}\quad \subfigure[]{\includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig1b.ps}} \end{center} \caption{ a) Snapshot of a sample where the spins on one of the sub-lattices were frozen and the spins on the other sub-lattice evolved accordingly to \ref{def.eq}. b) The time dependence of the time correlation function $q(t)$ for the spins on the $B$ sub-lattice. We can see that the curve relaxes to a non-zero value in accordance with the RFIM expectations.} \label{fixed.fig} \end{figure} } This result shows that the $B$ domains do not behave as spins in a traditional spin-glass model, each spin points to a given preferential direction for all temperatures, in total disagreement with the traditional spin-glass model where the spin orientation is random at all temperatures except at $T=0$. Looking at the snapshot presented on figure \ref{fixed.fig} we can clearly see that the domain size is much bigger than $\sim 7$, the value proposed on reference \cite{parisi1998}. Sepp\"al\"a {\it et al} have made exact zero temperature configuration calculations for the RFIM \cite{seppala1998}. They define a ferromagnetic break up length scale $L_b$. For $L=L_b$ the RFIM ground state has a probability of $0.5$ to be purely ferromagnetic; for larger $L$ this probability decreases and the ground state magnetization tends to zero. However up to a critical value of the random field, $\Delta_c$, there will always be a percolating domain (whose weight tends to zero in the thermodynamic limit). $\Delta$ is defined as the root mean square random field in units of the ferromagnetic interaction $J$. For our model with the $A$ sub-lattice frozen, $\Delta = 2 \lambda$. From the data presented in \cite{seppala1998}, we can estimate that for $\lambda = 0.5, 0.7$ and 1.5 (the three cases we will discuss below), $L_b \sim 45, 22$ and $5$. For total lattice size $L$ the sub-lattice size is $L/\sqrt{2}$, so that for samples of size $L=64$ the sub-lattices are close to, above, and well above $L_b$ respectively for these three $\lambda$ values. $\Delta_c$ corresponds to $\lambda = 1$. For the ground states of the present model with both sub-lattices free, the value of $L_b$ may not be quite the same as for the pure RFIM. However for the $\lambda=0.5$ data we will discuss below, we can expect many of the exact ground states to have complete sub-lattice ferromagnetic ordering at zero temperature, except for the largest sizes. At finite temperatures however the spin configurations in thermodynamic equilibrium will have domains that are large but smaller than the ground state domains. Now turn to the full model with non-zero $\lambda$ where both sub-lattices are free. For high enough temperatures the ferromagnetic domains on each sub-lattice are unstable. At low temperatures the $A$ and $B$ sub-lattices will conspire so that each induces random fields on the other such that the total energy is minimized. We will show evidence below that for $\lambda$ up to about 1 there is a low temperature state with frozen large sub-lattice ferromagnetic domains. For higher $\lambda$ the system appears to be paramagnetic at all temperatures, like the standard 2d ISG. \section*{Criteria for an ordering temperature} There are a number of different criteria which have been used in numerical work to determine the value of ordering temperatures in spin glasses and other complex Ising systems. Finite size scaling on the ``spin glass susceptibility'' is one of these. The spin glass susceptibility is defined by \begin{equation} \chi=L^d \langle q^2 \rangle \end{equation} where $ \langle q^2 \rangle $ represents the second moment of the equilibrium autocorrelation function fluctuations and $\langle \ldots \rangle$ represents both a configurational and thermal average. If corrections to finite size scaling are negligible, the spin glass susceptibility follows a scaling rule \cite{bhatt1988} \begin{equation} \chi(L,T)=L^{d-2+\eta}\, \mbox{f}\,(L^{1/\nu} (T-T_g)), \end{equation} meaning that precisely at $T_g$, $\log (\chi(L,T_g))$ plotted against $\log(L)$ should give a straight line of slope $d-2+\eta$. At higher temperatures $\chi (L)$ should saturate with increasing $L$ while at temperatures below $T_g$ the log-log plot should curve upwards. In fact this autocorrelation function susceptibility will show critical behavior in general at a critical temperature, including cases like ferromagnets which have standard order parameters. It is thus a parameter which can be used to identify a transition without the need to specify the exact nature of the transition. A complementary finite size scaling method was introduced by Binder \cite{binder1981}. The dimensionless Binder cumulant \begin{equation} g_L =\frac{1}{2}\left[ 3- \frac{\langle q^4\rangle}{ \langle q^2\rangle^2} \right] \end{equation} is a parameter characteristic of the shape of the distribution $P(q)$ of the equilibrium autocorrelation function fluctuations for a given sample of size $L$ and temperature $T$. For a system with a single characteristic length scale, $g_L$ is size independent at the ordering temperature, and a scaling rule applies: \begin{equation} g_L(T-T_c)=g(L^{1/\nu} (T-T_c)). \end{equation} Plots of $g_L(T)$ for different sizes $L$ should all intersect at $T_g$. This criterion has been widely used, particularly in the spin glass context. In this model it is also useful to consider the magnetization Binder parameter defined in the same way: \begin{equation} g_{mL} =\frac{1}{2}\left[ 3- \frac{\langle m^4\rangle}{ \langle m^2\rangle^2} \right]. \end{equation} The time dependence of the autocorrelation function $q(t)$ provides a further and fundamental criterion for an ordering temperature. In the thermodynamic limit, for the paramagnetic state in zero external field, $q(t)$ will always tend to zero at long $t$. As pointed out by Edwards and Anderson \cite{edwards1975}, if $q(t)$ tends to a finite long time limit the system can be considered to be ordered; the limiting value of $q(t)$ at temperature $T$ is the Edwards-Anderson ordering parameter. For a continuous phase transition, the time scale characteristic of the decay of $q(t)$ will diverge as the critical temperature is approached from above. This criterion was used to estimate $T_g$ accurately in the 3 dimension Ising spin glass measurements of Ogielski \cite{ogielski1985}, who assumed a standard finite ordering temperature scaling for the characteristic relaxation time $\tau(T)$. The $T_g$ value defined in this way has been confirmed later by independent finite size scaling methods. Marinari et al \cite{marinari1994} found that for the 3d Ising spin glass the temperature dependence of the spin glass susceptibility could be fitted equally well by 3 parameter expressions corresponding either to finite temperature or to zero temperature ordering. It appears that the temperature dependence of the relaxation time is much more discriminating than that of the temperature dependence of the susceptibility. \section*{Simulation techniques and Data} In order to answer the question of whether the freezing temperature is finite or not, we have made further simulations for $\lambda=0.5, 0.7$ and 1.5. Wherever direct comparisons could be made, our data are in full agreement with those of Parisi et al \cite{parisi1998}. Simulations were carried out using sequential heat bath updating. Samples up to size $64^2$ were studied. Table \ref{samples.tab} present the maximum annealing time and the number of different realizations for each size. The criterion used to determine if thermal equilibrium had been attained was the saturation of the two replica overlap as a function of anneal time \cite{bhatt1988}. \ifthenelse{\equal{condmat}{condmat}}{ \begin{table} \begin{tabular}{| c | c | c | } L & Samples & Anneal Time \\ \hline 4 & 10000 & 15000 \\ 8 & 10000 & 15000 \\ 16 & 1000 & 150000 \\ 32 & 500 & 150000 \\ 64 & 500 & 150000 \\ \end{tabular} \caption{The anneal times and the number of different realizations for each size studied.} \label{samples.tab} \end{table} } \subsection{Susceptibility} For convenience we have followed Parisi et al who used a non-standard spin glass susceptibility defined by \begin{equation} \chi_P =L^2\left[\langle q^2 \rangle - \langle |q^2 | \rangle \right] \end{equation} (c.f. the standard spin glass susceptibility defined above) In figure \ref{xi.fig} we present the dependence of the susceptibility $\chi_P$ with the size $L$ for different temperatures and for $\lambda=0.5$. The figure shows that the data follow precisely the behavior to be expected for a system with an ordering temperature lying somewhere between $2.1$ and $2.2$. For higher temperatures the susceptibility saturates with increasing size; for lower temperatures the $\log(\chi_P)$ against $\log(L)$ plot curves upwards. At around $T=2.2$, $\log(\chi_P)$ increases linearly with $\log(L)$, which is the signature of critical behaviour. From the slope of the critical line, the exponent $\eta$ can be estimated to be 1.2$\pm$0.1 . \ifthenelse{\equal{condmat}{condmat}}{ \begin{figure} \begin{center} \includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig2.ps} \end{center} \caption{ The dependence of the spin glass susceptibility with $L$ for different temperatures. For $T\sim 2.2$ we have $\chi\sim L^\alpha$. } \label{xi.fig} \end{figure} } \subsection{Binder Cumulant} In \cite{lemke1996b} the Binder cumulant crossing method for sample sizes up to $L=12$ was used to estimate ordering. Parisi et al \cite{parisi1998} showed that for larger sample sizes the crossing point of the Binder curves $g_L(T)$ moved to lower temperatures and became badly defined. On the basis of this observation they suggested that in fact for large sizes there is no ordering temperature, and that the low temperature state is paramagnetic. The Binder cumulant method can be delicate to use. This can be illustrated by a trivial ``paradox'' in the present system. For any standard Ising ferromagnet, $g_L(T)$ goes to $1$ at low temperatures. However in the present system, if $\lambda$ is zero (so each of the two sub-lattices order ferromagnetically), because of the four possible ground states $g_L(T)$ goes to $0.5$ at low temperatures, not to $1$. In the general case, $g_L(T)$ curves going to small values or even to zero at low temperatures for large systems is not the signature of a paramagnetic state, but rather of the system having a large number of orthogonal ground states. The classical Binder cumulant behavior with a well defined crossing point and good scaling above and below the critical temperature will be observed for systems with a single effective correlation length and a standard evolution with size and temperature for the form of the distribution $P(q)$. Certain non-standard systems with {\it bona fide} ordering transitions show very unorthodox behavior for the Binder cumulants \cite{bernardi1999}. Instead of concentrating attention on the details of the Binder cumulant curves at the lowest temperatures, we can do trial scaling plots over the whole temperature range for $\lambda=0.5$. For the scaling plots we can assume: \begin{itemize} \item that there is a critical temperature at about $T=2.1$ as indicated by the susceptibility scaling, or \item that there is a zero temperature critical point and an exponent $\nu$ equal to the value obtained from scaling of the Binder cumulant for data on the standard 2d ISG, i.e. $\nu=1.4\pm 0.2$ \cite{bhatt1988}. \end{itemize} The two scaling plots are shown in figure \ref{scale.fig}, while the raw data is presented on figure \ref{gl.fig}. It can be seen immediately that the first assumption leads to acceptable scaling, at least above the assumed critical point. The zero temperature scaling is clearly incorrect. We conclude that the finite critical temperature scaling is compatible with the Binder cumulant data while the zero temperature scaling is not. \ifthenelse{\equal{condmat}{condmat}}{ \begin{figure} \begin{center} \subfigure[]{\includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig3a.ps}}\quad \subfigure[]{\includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig3b.ps}} \end{center} \caption{ In figure a) the scaling proposed by reference Parisi et al in b) the scaling obtained by supposing $T_g=2.05$ and $\nu=1.4$. The data for $L=48$ were obtained from Parisi {\it et al}.} \label{scale.fig} \end{figure} \begin{figure} \begin{center} \includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig4.ps} \end{center} \caption{ The dependence of the Binder parameter with $L$ for different temperatures.} \label{gl.fig} \end{figure} } The fact, underlined by Parisi et al, that the low temperature Binder cumulant values do not increase regularly with increasing $L$ does not indicate that the low temperature state is paramagnetic, but that the type of order at low temperatures is evolving as $L$ increases. Direct evaluation shows that for $L=4$ there are frequently just two ground states, a purely ferromagnetic or antiferromagnetic state and its mirror image. This ``staggered ferromagnetism'' is what would be expected from the discussion of the RFIM given above; the sub-lattice magnetizations are essentially ferromagnetic and in any particular sample the random interactions select a ferro or an anti-ferro coupling between sub-lattices. In consequence for these small samples $g_L$ must tend to exactly $1$ at zero temperature and will be close to 1 at higher temperatures. This behavior will continue to hold until sizes are reached where $L$ is of the order of the $L_b$ at the particular temperature studied. If for larger samples there are many alternative more complex Gibbs states at that temperature, $g_L(T)$ will become lower for larger $L$. This seems to be the real situation, with temperature dependent crossover sizes. The sub-lattice magnetism Binder cumulant tends to $\sim 1$ at temperature$T \sim 1.5$ for samples up to size $L=8$, and then decreases regularly with increasing sample size \cite{parisi1998} figure 4. This indicates that the sub-lattices are ferromagnetic in the small samples, and in the larger samples each sub-lattice is principally either up or down (not zero magnetization as for large samples in the strict RFIM) but contain domains of non majority spin, at least at finite temperatures. As $L$ increases the average sub-lattice magnetization drops, but it would need very large $L$ for the sub-lattice magnetization distribution to take up a Gaussian form centered on zero \cite{seppala1998}. This gradual evolution with sample size, most clearly observed in the sub-lattice magnetization cumulant, is certainly also the cause of the ``anomalous'' low temperature behaviour of the sublattice $g_L$ cumulant and the global $g_L$ cumulant at low temperatures (see figures 7 and 9 of reference \cite{parisi1998}). From the discussion above, in these particular systems we can expect deviations from asymptotic large scale behaviour until very large values of $L$, well beyond the values used so far in the simulations. Finally, it must be remembered that a Binder cumulant is not directly sensitive to whether the spins are frozen or not. \subsection{Relaxation} We measured the autocorrelation function decay $q(t)$ after long equilibration anneals at different temperatures for large samples, $L=64$. This was done for $\lambda =0.5, 0.7$, and $1.5$; the results are presented on figures \ref{qA.fig}, \ref{qB.fig} and \ref{qC.fig}. At each temperature the form of the decay can be seen to be initially algebraic $q(t) \sim t^{-x}$, with a cutoff function at longer times. For the first two values of $\lambda$, as $T$ is reduced towards a temperature close to 2.1, the relaxation becomes purely algebraic to long time scales, meaning that the characteristic time defining the cutoff function is diverging. The characteristic time for the decay can be defined either by $\tau_c$ or by $\tau_{av}$ where \begin{eqnarray} \tau_c &=& \int_0^\infty q(t), \\ \tau_{av} &=& \frac{\int_0^\infty tq(t)} {\int_0^\infty q(t)} \\ \end{eqnarray} $\tau_c$ and $\tau_{av}$ were calculated for convenience by fitting the $q(t)$ curves with an Ogielski function \cite{ogielski1985} : \begin{equation} q(t)=t^{-x}\exp\left[ -\left(\frac{t}{\tau} \right)^\beta\right] . \end{equation} \ifthenelse{\equal{condmat}{condmat}}{ \begin{figure} \begin{center} \includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig5.ps} \end{center} \caption{ The time dependence of the correlation function for $\lambda=0.5$ and different temperatures. From bottom to top $T=$ 3.5, 3.0, 2.7, 2.6, 2.5, 2.4, 2.3, 2.2.} \label{qA.fig} \end{figure} \begin{figure} \begin{center} \includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig6.ps} \end{center} \caption{ The time dependence of the correlation function for $\lambda=0.7$ and different temperatures. From bottom to top $T=$ 3.5, 3.0, 2.5, 2.3, 2.2, 2.0.} \label{qB.fig} \end{figure} \begin{figure} \begin{center} \includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig7.ps} \end{center} \caption{ The time dependence of the correlation function for $\lambda=1.5$ and different temperatures. From bottom to top $T=$ 3.5, 3.0, 2.5, 2.0.} \label{qC.fig} \end{figure} } For $\lambda =0.5$, $\tau_c(T)$ and $\tau_{av}(T)$ diverge at a temperature just below $T=2.1$, figure \ref{tau.fig}. The critical value of the exponent $x$ at the temperature where $\tau(T)$ diverges is about 0.15. For $\lambda =0.7$ the behaviour is very similar and the exponent is about 0.11. This means that for both these values of $\lambda$, below a critical temperature and in the thermodynamic limit the system is frozen to indefinitely long times with a finite Edwards Anderson parameter, and so it is ordered (in the Edwards Anderson sense). We can note that the behaviour is very similar for these two values of $\lambda$, although from the discussion above we can expect the latter to have a break up length $L_b$ half as large as in the former.For $\lambda = 0.5$ the zero temperature $L_b$ estimated above is about equal to the sublattice size (which we can take equal to $L/\sqrt(2)$) while for $\lambda =0.7$ $L_b$ is significantly smaller than the lattice size. This criterion does not appear to play a major role at the temperature where the relaxation time is diverging. \ifthenelse{\equal{condmat}{condmat}}{ \begin{figure} \begin{center} \includegraphics[bb= 0 0 600 800, scale=0.3, angle= 270]{fig8.ps} \end{center} \caption{ The temperature dependence of $\tau_c$ with temperature for $\lambda$ =0.5, 0.7, 1.5. The figure shows clearly that $\tau_c$ is diverging for $\lambda$ =0.5 and 0.7, and grows slowly for $\lambda$ =1.5. } \label{tau.fig} \end{figure} } The behaviour for $\lambda=1.5$ is quite different; the temperature variation of the relaxation time is very much slower, figure \ref{qC.fig} and \ref{tau.fig}. In this case the form of $\tau (T)$ can be compared with that seen in the standard 2d ISG case, where $\tau(T)$ is seen to diverge only as zero temperature is approached, following an Arrhenius like law \cite{mcmillan1983}. The $q(t)$ measurements are clearly discriminatory and can distinguish systems with finite freezing temperatures from ones with zero temperature (or at least very low temperature) ordering. The susceptibility scaling, Binder cumulant scaling, and autocorrelation relaxation data thus give conclusive and consistent evidence for freezing at a temperature near 2.1 for $\lambda=0.5$. The relaxation data indicate a slightly lower freezing temperature for $\lambda =0.7$, and a much lower temperature freezing compatible with $T_g=0$ for $\lambda= 1.5$. The estimated freezing temperatures are very similar to those suggested originally in \cite{lemke1996b}. There seems to be no evidence for an onset of paramagnetic behaviour at low temperatures with increasing size. We conclude that the low temperature state is frozen for $\lambda$ values up to about 1. It is perhaps not a coincidence that the critical value of $\lambda$ appears to correspond to the critical RFIM $\Delta_c$ as defined above. Can the present system be described as a spin glass ? We can attempt to give a coherent description of the low temperature frozen state in the light of the different types of data. As we have seen, for small sizes the ordering can indeed be described in terms of ``staggered ferromagnetism''. For the larger sizes covered in this work and in \cite{parisi1998}, the low temperature sublattice magnetism Binder cumulant decreases regularly with increasing $L$, but is still as high as 0.8 at $L=48$ and $T=1.5$. (\cite{parisi1998} figure 4). This indicates that in equilibrium at this temperature, each sublattice is split up into fairly big ferromagnetic domains with magnetization of both signs, but for each particular replica, one sign of magnetization is preponderant for each sublattice. However within each sublattice there are large minority domains. We have found that if a large sample is cooled a number of different times to a temperature below the critical temperature estimated above, the quasi-stationary pattern of domains observed is far from identical each time (in contrast to what is always seen in the standard RFIM). We can extrapolate, and surmise that for very large $L$ there would be no magnetization bias for a sublattice, and there would then be for the entire system a very large number of possible Gibbs states below the ordering temperature, nearly orthogonal to each other in phase space. The whole system can be understood as freezing at low temperatures because once domains of a maximal size are formed, the domain walls are pinned by the effect of the random interactions. The low temperature state would then ressemble a spin glass in that there are many Gibbs states, but the local spin structure is entirely different because of the strong local ferromagnetic correlations within each sublattice. \section*{Conclusions} We have investigated in more detail the ferromagnetic plus random interaction system described by equation (\ref{def.eq}). To summarize: data on autocorrelation function or ``spin glass'' susceptibility, Binder cumulant, and autocorrelation function relaxation, all consistently indicate a critical temperature for freezing of $T_c \sim 2.1$ for $\lambda =0.5$. Relaxation data indicate a slightly lower freezing temperature for $\lambda =0.7$, and are compatible with a zero temperature freezing for $\lambda =1.5$. Therefore in the range of $\lambda$ up to about 1, this two dimensional RCF system with interactions which are partly random has a finite freezing temperature. There is no evidence for a return to paramagnetic behaviour (with faster relaxation for instance) with increasing size. Independent defect free energy data confirm our conclusions \cite{shirakura1998}. The finite freezing temperature result is not in contradiction with the general consensus that for standard 2D spin glass model the ordering temperatures are either zero or at least very low. The picture for the transition suggested above implies that the transition mechanism for this system is radically different from that of the standard spin glass. We have no reason to expect that this system can be mapped onto a standard Ising spin-glass, even though this system shares many properties with the traditional model: frustration, complex phase space landscape etc. Because of the ferromagnetic short range ordering within each sublattice, the term ``cluster glass'' would probably be more appropriate than ``spin glass''. Many interesting questions remain; in particular it will be very important to describe accurately the nature of the low temperature phase, and to obtain explicity information about the domain size distribution, and the domain geometry characteristics in large samples and at low temperatures. It should be possible to apply sophisticated methods to establish ground state characteristics. Finally we believe this model should be a very useful laboratory to test theoretical issues concerning disordered systems, since in this case we have a freezing transition in a two-dimensional system, where theoretical analysis, exact ground state methods, simulations, and visualization techniques are easier to apply than in higher dimensions. Further progress would however require the study of much larger samples. \acknowledgments We would like to thank T. Shirakura for showing us his unpublished data. N. L. would like to thank the kind hospitality of the Laboratoire de Physique des Solides during the preparation of this manuscript. \bibliographystyle{prsty}
\section{Introduction} Suppose we are given a symplectic manifold $(P,\omega)$ and a compact Lie group $G$ acting freely and symplectically on $P$ with $\mbox{Ad}^*$-equivariant momentum mapping $J:P\rightarrow\mathfrak g^*$. Let the Lie algebra of $G$ be $\mathfrak g$ and $H:P\rightarrow\mathbb R$ be a $G$-invariant Hamiltonian. Let us fix our attention on a specific relative equilibrium~$p_e$, with generator~$\xi_e$, so that the integral curve of the Hamiltonian vector field $X_H$ at $p_e$ is $\exp({\xi_e} t)p_e$. Let~$J(p_e)={\mu_e}$, let $G_{\mu_e}$ be the isotropy group of ${\mu_e}$ under the coadjoint action of $G$ on $\mathfrak g^*$, and let $\mathfrak g_{\mu_e}$ be the Lie algebra of $G_{\mu_e}$. For more information on these basic definitions, see, for example,~\ct{AbrahamRMarsdenJE-1978.1},~\ct{MarsdenJE-1992.1}, or~\ct{MarsdenJE-1994.1}. Suppose that $G=SO(3)$, although the results of this article are much more general than that. Since the action is free, a neighborhood of the group orbit $Gp_e$ may be (with respect to the left action) equivariantly projected to $SO(3)$ itself. Let us realize $SO(3)$ as the unit circle bundle of the 2-sphere in $\mathbb R^3$, arranging that the orbit $\exp(\xi t)p_e$ be projected to the the unit circle in the tangent space of the vertical vector $\mathbf k$. The system, when at the relative equilibrium $p_e$, and when viewed through this projection, appears merely as a point, say $\mathcal P$, rotating uniformly on the unit circle in the tangent space of $S^2$ at $\mathbf k$, and the other relative equilibria $gp_e$, $g\in SO(3)$, appear as this same motion but reoriented by the rotation $g$. Suppose $p_e$ is formally stable, and suppose the system is then perturbed from the relative equilibrium $p_e$. If $\mu_e\ne 0$ then the motion of the system is bound to remain near $\exp(\xi_e\mathbb R)p_e$, as shown in~\ct{PatrickGW-1992.1} (extensions to the nonfree case may be found in~\ct{OrtegaJPRatiuTS-1997.1} and~\ct{LermanESingerSF-1997.1}). The perturbation affects the motion of the point $\mathcal P$ by superimposing on its original motion some small vibration; however the motion of $\mathcal P$ does not ever carry that point far from its original circular path. The situation is different when ${\mu_e}=0$, for then $G_{{\mu_e}}=SO(3)$, and under perturbation the system is bound only to remain near $Gp_e$. The effect is that, after a perturbation from $p_e$, the point $\mathcal P$ still moves nearly circularly in tangent spaces of $S^2$, but the base point of the tangent space slowly changes location on $S^2$. In fact, as shown in~\ct{PatrickGW-1995.1}, the base point $\mathcal B$ moves as a point charge under the influence of a radial magnetic monopole. This result is obtained by constructing coordinates which make an open neighborhood of $Gp_e$ in $P$ into an open neighborhood of $Z\bigl(T^*SO(3)\bigr)\times \{\bar p_e\}$ in $T^*SO(3)\times P_{\mu_e}$, where $P_{\mu_e}=J^{-1}({\mu_e})/G_{{\mu_e}}$ is the symplectic reduced space, $\bar p_e\in P_{\mu_e}$ is the equilibrium of the reduced space corresponding to $p_e$, and $Z\bigl(T^*SO(3)\bigr)$ is the zero section of $T^*SO$. In the new coordinates, to first order in the momentum perturbation, the ``group variables'' in $T^*SO(3)$ and the ``reduced variables'' in $P_{\mu_e}$ become decoupled; truncating the higher order interaction terms gives a new Hamiltonian system on $T^*SO(3)\times P_{\mu_e}$ called the \defemph{drift system}. The drift system has an additional $S^1$ \defemph{normal form symmetry}. The Hamiltonian system for the slow motion of the base point $\mathcal B$ is obtained from the drift system by ignoring the variables on $P_{\mu_e}$ and reducing the resulting Hamiltonian on $T^*SO(3)$ by the normal form symmetry. The approximating, truncated Hamiltonian of the drift system may be obtained merely by calculating the nilpotent part of the linearization of the vector field $X_{H_{\xi_e}}(p)\equiv X_H(p)-\xi_ep$ at $p_e$, where $\xi p\equiv (\xi)_P(p)$ denotes the infinitesimal generator of $\xi_e$ at $p_e$. Actually, the results in~\ct{PatrickGW-1995.1} depend on a nonresonance condition that the frequency $|\xi_e|$ cannot equal any linearized frequency of the reduced system at $\bar p_e$. This article considers the problem of constructing a Hamiltonian system for the motion of the base point $\mathcal B$ in the case where there is such a \defemph{1:1 group-reduced resonance}. The main result of this article is that, in the presence of resonance, some of the reduced and group variables couple to first order in the momentum perturbation, with the result that $\mathcal B$ moves as a charged particle~\emph{with magnetic moment} on a sphere under the influence of a radial magnetic monopole. However, this ``particle'' regularly exchanges its charge with its magnetic moment. The construction of the Hamiltonian system which models the motion of $\mathcal B$ in the 1:1 resonant case generalizes and parallels the nonresonant case. In the course of the construction the linearization of $X_{H_{\xi_e}}$ at $p_e$ is put into a certain block normal form through a Witt or Moncreif decomposition of the tangent space $T_{p_e}P$. The objective of this normal form is the separation of the group and reduced motions and it is not the full normal form of the linearization as a infinitesimal symplectic operator. The linear normal form is then extended to a neighborhood of the group orbit $Gp_e$ using the the isotropic embedding/equivariant Darboux theorem. These normal forms---for linearizations of relative equilibria and for Hamiltonian systems near group orbits of relative equilibria---are the second main contribution of this article. In~\ct{PatrickGWRobertsRM-1999.1} these normal forms have been used to study the structure of the set of relative equilibria. A new aspect that is absent in the nonresonant case but emerges in the resonant case is the appearance of a~\defemph{gauge group}. There is too much freedom inherent in the normal forms to exactly fit the needs of first order agreement between the original and drift systems. This slack is taken up by the group $(\mathbb R,+)$ and presents itself as an inherent freedom in the choice of the normal forms. This~\emph{gauge freedom} can be used to simplify the drift system, a motif which is quite useful in this work. The Hamiltonian system modeling the motion of $\mathcal B$ is approximate since it is the result of a truncation, so there arises the question of how well this model Hamiltonian reflects the behavior of the real one. To check this I have numerically simulated the system of two (identical) axially symmetric rods which are joined by a frictionless ball-and-socket joint, but otherwise move freely~(\ctt{PatrickGW-1989.1}{PatrickGW-1991.1}). The relative equilibria of this system correspond to motions such that the two rods spin on their axes while otherwise maintaining constant mutual orientations, while the whole assemblage rotates about some fixed axis. There exist, due to the system's multiple rotating parts, relative equilibria which are formally stable and which have zero (hence nongeneric) total-angular-momentum. In fact there is a continuum of relative equilibria with zero total-angular-momentum, and it happens that parameters may be chosen so that there are relative equilibria at zero total-angular-momentum and at the 1:1 group-reduced resonance. The third main contribution of this article is the actual calculation of the drift system for the coupled rod system near one such relative equilibria, and the verification of the drift approximation by comparison of simulations of the coupled rod system itself and certain predictions of its drift system. Here is an overview of this work. I begin in Section~2.1 with an analysis of the linearization~$dX_{H_{\xi_e}}(p_e)$ of $X_{H_{\xi_e}}$ at $p_e$, by splitting the linearization into semisimple and nilpotent parts: $dX_{H_{\xi_e}}(p_e)= S_{p_e}+ N_{p_e}$. The key tool in the analysis is a certain Moncrief or Witt decomposition of~$T_{p_e}P$: \begin{equation*} T_{p_e} P\congW_{\!\mathrm{red}}\oplus(\mathfrak g_{\mu_e}\oplus \mathfrak g_{\mu_e}^*)\oplus T_{\mu_e}(G{\mu_e}). \end{equation*} Whereas this decomposition is obtained in~\ct{PatrickGW-1995.1} for the nonresonant case with the aid of the generalized eigenspaces of $dX_{H_{\xi_e}}(p_e)$, and so was $dX_{H_{\xi_e}}(p_e)$-invariant, here the decomposition is only $S_{p_e}$-invariant. Using this decomposition, the nilpotent part~$ N_{p_e}$ acquires a block form which gives rise to certain operators: \begin{equation*} N_{p_e}^{23}:\mathfrak g_{\mu_e}^*\rightarrow\mathfrak g_{\mu_e},\quad N_{p_e}^{21}:T_{\bar p_e}P_{\mu_e}\rightarrow\mathfrak g_{\mu_e}, \quad N_{p_e}^{13}:\mathfrak g_{\mu_e}^*\rightarrow T_{\bar p_e}P_{\mu_e}. \end{equation*} In the nonresonant case $N_{p_e}^{21}=0$ and $N_{p_e}^{13}=0$ necessarily. I expose, in Section~2.2, the special properties of these operators, the most notable of which are that $N_{p_e}^{23}$ is symmetric, $N_{p_e}^{13}$ and $N_{p_e}^{21}$ are dual, and that these operators have certain commutation relations with operators $\onm{ad}_{\xi_e}$ and $\onm{coad}_{\xi_e}$. Next, in Section~3, I focus on the case where the reduced and group spectra have intersection $\{\pm i\lambda_{p_e}\}$, a purely imaginary eigenvalue and its conjugate, each of which occur with multiplicity~1 in both the reduced and group spectrum. The generalized eigenspace of the reduced linearization corresponding to $\{\pm i\lambda_{p_e}\}$ is symplectic and has dimension~2, and so is linearly symplectomorphic to $\bigl(\mathbb R^2=\{x_1,x_2\},dx_1\wedge dx_2\bigr)$, and the operator $N_{p_e}^{21}$ is zero on the sum of the complimentary generalized eigenspaces. Consequently, the splitting of $T_{p_e}P$ refines and the operator $N_{p_e}^{21}$ may be replaced by another operator $N_{p_e}^{211}:\mathbb R^2\rightarrow\mathfrak g_{\mu_e}$. After this simplification, the splitting of $T_{p_e}P$ and the isotropic embedding theorem~(\ct{MarsdenJE-1981.1},\ct{WeinsteinA-1997.1}) together give a map which transforms the Hamiltonian $H$, to first order near $G_{\mu_e} p_e$, to the Hamiltonian \begin{equation*} H_{\mathrm{drift}}(x,\alpha_g)\equiv\langle g^{-1}\alpha_g,\xi_e\rangle+ \frac12N_{p_e}^{23}(\alpha_g,\alpha_g)+ \langle g^{-1}\alpha_g, N_{p_e}^{211}x\rangle \end{equation*} near $\{0\}\times Z(T^*G_{\mu_e})$ in the phase space $\mathbb R^2\times T^*G_{\mu_e}$, where $ Z(T^*G_{\mu_e})$ denotes the zero section. In addition to the expected invariance under the left action of $G_{\mu_e}$, $H_{\mathrm{drift}}$ is invariant under a diagonal action of the toral subgroup generated by $\xi_e$; this is the normal form symmetry. In Section~4 I further assume that $G$ is the largest compact continuous symmetry group for an ordinary mechanical system, namely $SO(3)\times(S^1)^n$. After various manipulations, including an Abelian reduction, the drift Hamiltonian becomes \begin{equation*} H_{\mathrm{drift}}\equiv\frac12I_1\bigl({\pi_1}^2+{\pi_2}^2\bigr)+\frac12I_2{\pi_3}^2+ \kappa(\pi_1x_1+\pi_2x_2)+a\pi_3 \end{equation*} on the phase space $\mathbb R^2\times T^*SO(3)=\{(x,A,\pi)\}$. Here $I_1$ and $I_2$ come from $N_{p_e}^{23}$, $\kappa$, a single coupling constant, is all that remains of $N_{p_e}^{211}$, $a$ is a constant, and there are two possibilities: the relative equilibrium can be either a ``+'' type or a ``$-$'' type. In this context the gauge freedom has the effect of making $I_1$ arbitrary and interest is focused on the dynamics near to $x=0$, $\pi=0$. The Hamiltonian $H_{\mathrm{drift}}$ is defined on a phase space of dimension~$8$ and has symmetry $SO(3)\times S^1$, and so defines a system which is completely integrable. Although a complete analysis in the general case seems difficult, I analyze this system in Section~4.1, where I show that the system has, for example, some singular reduced phase spaces and spectrally unstable relative equilibria with homoclinic connections. It is in Section~4.2 that I show that the drift system reduced by the normal form symmetry can be cast as a charged particle with magnetic moment moving on the sphere while under the influence of a magnetic monopole. Finally, in Section~5, I numerically investigate the dynamics of the coupled rod system near one particular resonant relative equilibrium and compare this dynamics with that of the drift system. I explore three distinct regions of phase space: 1) within zero total-angular-momentum, 2) near a stable relative equilibria of the drift system, and 3) near a spectrally unstable relative equilibrium at a singularity of the drift system. The comparison of the two systems is hampered by the implicit nature of the coordinates relating them, but in the first two comparisons agreement between the two systems is obtained uneventfully. However, the singular points of the third comparison involve an unexpected \emph{reconstuction phase jump}; the situation is delicate and small perturbations are required to elicit quantitative agreement between the two systems. Nevertheless it becomes clear that many elements of the dynamics of the coupled rod system near the resonant relative equilibrium are indeed captured by the drift system. \section{The linearization} Here are the basic notations: \begin{enumerate} \item $p_e$ is a relative equilibrium, $\xi_e$ is the generator of $p_e$, and the momentum of $p_e$ is ${\mu_e}\equiv J(p_e)$. \item $G_{\mu_e}$ is the isotropy group of ${\mu_e}$ under the coadjoint action of $G$ on $\mathfrak g^*$, and $\mathfrak g_{\mu_e}$ is the Lie algebra of $G_{\mu_e}$. $\onm{CoAd_g}\equiv(\onm{Ad}_{g^{-1}})^*$, $g\in G$, and $\onm{coad}_\xi\equiv-(\onm{ad}_\xi)^*$, $\xi\in\mathfrak g$. \item $H_{\xi_e}\equiv H-J_{\xi_e}$, so that $H_{\xi_e}$ has a critical point at $p_e$. The Hessian of ${H_{\xi_e}}$ at $p_e$ will be denoted by $d^2{H_{\xi_e}}(p_e)$ and the linearization of $X_{H_{\xi_e}}$ at $p_e$ will be denoted $dX_{H_{\xi_e}}(p_e)$. \item Without loss of generality, since everything is local near to $p_e$ and $p_e$ is regular, the Marsden-Weinstein symplectic reduced space $(P_{\mu_e},\omega_{\mu_e})$ exists; let $\pi_{\mu_e}:J^{-1}({\mu_e})\rightarrow P_{\mu_e}$ be the projection. The reduced system has this phase space with Hamiltonian $H_{\mu_e}$ defined by $H_{\mu_e}\pi_{\mu_e}=H|J^{-1}({\mu_e})$. Also, $\bar p_e\equiv\pi_{\mu_e}(p_e)$, and the linearization of the reduced system at $\bar p_e$ is denoted by $dX_{H_{\mu_e}}(\bar p_e)$. Suppose that $p_e$ is regular, which means $\xi p_e\neq0$ for all $\xi\in\mathfrak g$. \item The \defemph{reduced spectrum} is the spectrum of the linearization $dX_{H_{\mu_e}}(\bar p_e)$ at $\bar p_e$ of the reduced system. The \defemph{group spectrum} is the the spectrum of $\onm{ad}_{\xi_e}:\mathfrak g\rightarrow\mathfrak g$). \end{enumerate} Let $dX_{H_{\xi_e}}(p_e)= N_{p_e}+ S_{p_e}$ be the Jordan decomposition of $dX_{H_{\xi_e}}(p_e)$ into its semisimple part $S_{p_e}$ and nilpotent part $ N_{p_e}$. The aim of this section is an analysis of $dX_{H_{\xi_e}}(p_e)$, focusing on its nilpotent part $N_{p_e}$. \emph{Do not} assume that the reduced spectrum and the group spectrum are disjoint. \subsection{The splitting of $T_{p_e}P$} I begin by deriving a ``Moncreif'' splitting~(\ct{MarsdenJE-1981.1}) of $T_{p_e}P$ which is slightly weaker than its analogue in~\ct{PatrickGW-1995.1}, in that the ``reduced'' part (below $W_{\!\mathrm{red}}$) of the splitting is not (and cannot, in general) be constructed to be $dX_{H_{\xi_e}}(p_e)$ invariant. The details are similar to those in~\ct{PatrickGW-1995.1}; I will not belabor them here. The subspace $\onm{ker}dJ(p_e)$ is $S_{p_e}$-invariant, so one can choose an $S_{p_e}$-invariant subspace $W_{\!\mathrm{red}}$ such that \begin{equation*}\onm{ker}dJ(p_e)=W_{\!\mathrm{red}}\oplus\mathfrak g_{\mu_e} p_e. \end{equation*} The subspace $W_{\!\mathrm{red}}$ is symplectic, and $T\pi_{\mu_e}|W_{\!\mathrm{red}}:W_{\!\mathrm{red}}\rightarrow T_{\bar p_e}P_{\mu_e}$ is a linear symplectomorphism. Choose an $\onm{Ad}$-invariant complement $\mathfrak b$ to $\mathfrak g_{\mu_e}$, so that \begin{equation*} \mathfrak g=\mathfrak b\oplus\mathfrak g_{\mu_e}. \end{equation*} The subspace $\mathfrak bp_e$ is symplectic, and $\mathfrak g_{\mu_e}\subseteq(W_{\!\mathrm{red}}\oplus\mathfrak bp_e)^{\omega\perp}$ as an $S_{p_e}$-invariant \emph{Lagrangian} subspace. Choose an $S_{p_e}$-invariant Lagrangian complement\footnote{Since $S_{p_e}$ is semisimple, every $S_{p_e}$-invariant Lagrangian subspace has an $S_{p_e}$-invariant Lagrangian complement. This is a general fact; the proof is a simple modification of the argument found at the bottom of page~401 of~\ct{PatrickGW-1995.1}.} $Z$ to $\mathfrak g_{\mu_e}$ in $(W_{\!\mathrm{red}}\oplus\mathfrak bp_e)^{\omega\perp}$, giving the $S_{p_e}$-invariant splitting \begin{equation}\lb{1} T_{p_e}P=W_{\!\mathrm{red}}\oplus\mathfrak g_{\mu_e} p_e\oplus Z\oplus\mathfrak bp_e. \end{equation} As already stated, $T\pi_{\mu_e}$ is a symplectomorphism between $W_{\!\mathrm{red}}$ and $T_{p_e}P_{\mu_e}$. Also, $dJ(p_e)$ is a linear isomorphism between $Z$ and $\mathfrak g_{\mu_e}^*$, and a symplectomorphism between $\mathfrak b p_e$ and the tangent space $T_{\mu_e}(G{\mu_e})$ at ${\mu_e}$ to the coadjoint orbit $G{\mu_e}$. Thus~\rf{1} becomes, through these identifications, the $ S_{p_e}$-invariant splitting, \begin{equation}\lb{2} T_{p_e}P=T_{\bar p_e}P_{\mu_e}\oplus\mathfrak g_{\mu_e}\oplus \mathfrak g_{\mu_e}^* \oplus T_{\mu_e}(G{\mu_e}). \end{equation} With respect to the decomposition~\rf{2}, let $dX_{H_{\xi_e}}(p_e)$, $S_{p_e}$, and $N_{p_e}$ have blocks $[A^{ij}]$, $[S^{ij}]$, and $[N_{p_e}^{ij}]$, respectively; the block form of $\omega(p_e)$ becomes \begin{equation}\lb{3} \omega(p_e)=\left[\begin{array}{cccc} \omega_{\mu_e}(\bar p_e)&0&0&0\\ 0&0&\Id&0\\ 0&-\Id&0&0\\ 0&0&0&\breve\omega_{\mu_e}({\mu_e}) \end{array}\right],\end{equation} where $\omega_{\mu_e}$ is the reduced symplectic form of $P_{\mu_e}$ and $\breve\omega_{\mu_e}$ is the Kostant-Souriau form on $G{\mu_e}$. In~\rf{1}, the subspaces $\mathfrak g_{\mu_e} p_e$, $\mathfrak b p_e$, and $\onm{ker}dJ(p_e)=W_{\!\mathrm{red}}\oplus\mathfrak g_{\mu_e} p_e$ are $dX_{H_{\xi_e}}(p_e)$-invariant, and hence are $N_{p_e}$-invariant, and this implies certain of the $A^{ij}$ and $N_{p_e}^{ij}$ vanish. Using the identities (Proposition~5 of~\ct{PatrickGW-1995.1}) \begin{gather*} X_{H_{\xi_e}}(p_e)=-\onm{coad}_{\xi_e}dJ(p_e)\\ dX_{H_{\xi_e}}(p_e)\eta(p_e)=-(\onm{ad}_{\xi_e}\eta)_P(p_e),\quad\eta\in \frak g \end{gather*} to calculate the diagonal blocks of $dX_{H_{\xi_e}}(p_e)$, the Jordan decomposition becomes \begin{equation*}\begin{split} dX_{H_{\xi_e}}(p_e)&=\left[\begin{array}{cccc} dX_{H_{\mu_e}}(\bar p_e)&0&A^{13}&0\\ \rule{0pt}{11pt}A^{21}&-\onm{ad}_{\xi_e}&A^{23}&0\\ 0&0&\onm{ad}^*_{\xi_e}&0\\ 0&0&0&\onm{ad}^*_{\xi_e} \end{array}\right]\\ &= \left[\begin{array}{cccc} S^{11}&0&0&0\\ 0&S^{22}&0&0\\ 0&0&S^{33}&0\\ 0&0&0&S^{44} \end{array}\right] +{\renewcommand\arraystretch{1.15}\left[\begin{array}{cccc} N_{p_e}^{11}&0&N_{p_e}^{13}&0\\ N_{p_e}^{21}&N_{p_e}^{22}&N_{p_e}^{23}&0\\ 0&0&N_{p_e}^{33}&0\\ 0&0&0&N_{p_e}^{44} \end{array}\right]}. \end{split}\end{equation*} Since $S_{p_e}$ and $N_{p_e}$ commute, so do $S^{11}$ and $N_{p_e}^{11}$, and $S^{11}$ is semisimple and $N_{p_e}^{11}$ is nilpotent since $S_{p_e}$ and $N_{p_e}$ are. Thus, $dX_{H_{\mu_e}}(\bar p_e)=S^{11}+N_{p_e}^{11}$ is the Jordan decomposition of of the reduced linearization $dX_{H_{\mu_e}}(\bar p_e)$, which is semisimple, since $p_e$ is formally stable. Consequently, $S^{11}=dX_{H_{\mu_e}}(\bar p_e)$ and $N_{p_e}^{11}=0$. Similarly, $N_{p_e}^{22}=N_{p_e}^{33}=N_{p_e}^{44}=0$, and \begin{equation*} S^{22}=-\onm{ad}_{\xi_e},\quad S^{33}=\onm{ad}^*_{\xi_e}, \quad S^{44}=\onm{ad}^*_{\xi_e}.\end{equation*} Thus, \begin{equation}\lb{4} S_{p_e}=\left[\begin{array}{cccc} dX_{H_{\mu_e}}(\bar p_e)&0&0&0\\ \rule{0pt}{11pt}0&-\onm{ad}_{\xi_e}&0&0\\ 0&0&(\onm{ad}_{\xi_e})^*&0\\ 0&0&0&(\onm{ad}_{\xi_e})^* \end{array}\right] \end{equation} and \begin{equation}\lb{5} N_{p_e}={\renewcommand\arraystretch{1.15}\left[\begin{array}{cccc} 0&0&N_{p_e}^{13}&0\\ N_{p_e}^{21}&0&N_{p_e}^{23}&0\\ 0&0&0&0\\ 0&0&0&0 \end{array}\right]}. \end{equation} As will be seen immediately below, a large amount of information can be discerned from structure of $N_{p_e}$ and $S_{p_e}$ visible in~\rf{4} and~\rf{5}. \subsection{Properties of $N_{p_e}$ and its subblocks} Directly from~\rf{5}, $( N_{p_e})^3=0$ so that $N_{p_e}$ has nilpotent order at most~3. This upper bound on the nilpotent order occurs because $dX_{H_{\xi_e}}(p_e)$ is semisimple when suitably restricted and projected to the summands of~\rf{1}, so that $N_{p_e}$ must ``transport'' between those summands. Moreover, the ``direction'' of this transport is ``one way'', since $\onm{ker}dJ(p_e)$ and $\mathfrak g_{\mu_e} p_e$ are invariant subspaces, and the ``transport'' stops at $\mathfrak g_{\mu_e} p_e$, on which $ N_{p_e}$ is zero: \begin{gather} \onm{Image} N_{p_e}\subseteq \onm{ker}dJ(p_e),\lb{6}\\ N_{p_e}\bigl(\onm{ker}dJ(p_e)\bigr)\subseteq\mathfrak g_{\mu_e} p_e,\nonumber\\ N_{p_e}|\mathfrak g_{\mu_e} p_e = 0.\lb{8} \end{gather} When iteratively acted upon by $N_{p_e}$, a vector can make just~2~stops before annihilation: $\onm{ker}dJ(p_e)$ and $\mathfrak g_{\mu_e} p_e$; hence the nilpotent order of $N_{p_e}$ is a most~3. By~\rf{8}, \begin{equation*} \onm{ker}T_{p_e}\pi_{\mu_e}=\mathfrak g_{\mu_e} p_e\subseteq\onm{ker} \bigl( N_{p_e}|\onm{ker}dJ(p_e)\bigr), \end{equation*} so $N_{p_e}^{21}:T_{\bar p_e}P_{\mu_e}\rightarrow\mathfrak g$ is the unique linear map such that \begin{equation}\lb{24}\begin{diagram} \node{\onm{ker}dJ(p_e)}\arrow{e,t}{N_{p_e}}\arrow{s,l}{T_{p_e}\!\pi_{\mu_e}} \node{\mathfrak g_{\mu_e} p_e}\arrow{e,t}{\xi p_e\mapsto\xi} \node{\mathfrak g_{\mu_e}}\\ \node{T_{\bar p_e}P_{\mu_e}}\arrow{ene,b}{N_{p_e}^{21}} \end{diagram}\end{equation} Since this diagram does not depend on choices made in the construction of the splitting~\rf{1}, neither does $N^{21}$. Also, $N_{p_e}^{21}$ may be calculated merely by reference to the diagram; it is not necessary to calculate all summands of~\rf{1}. Similarly, by~\rf{6}, and since, by~\rf{5}, \begin{equation*}\onm{ker}dJ(p_e)\subseteq\onm{ker}(T\pi_{\mu_e} N_{p_e}), \end{equation*} $N_{p_e}^{13}:\mathfrak g_{\mu_e}^*\rightarrow T_{\bar p_e}P_{\mu_e}$ is the unique linear map such that \begin{equation*}\begin{diagram} \node{T_{p_e}P}\arrow{e,t}{dJ(p_e)}\arrow{s,l}{N_{p_e}} \node{\frak g^*}\arrow{e}\node{\mathfrak g_{\mu_e}^*} \arrow[2]{sw,b}{N_{p_e}^{13}}\\ \node{\onm{ker}dJ(p_e)}\arrow{s,l}{T_{p_e}\!\pi_{\mu_e}}\\ \node{T_{\bar p_e}P_{\mu_e}} \end{diagram}\end{equation*} Since $dX_{H_{\xi_e}}(p_e)$ is infinitesimally symplectic, so is $N_{p_e}$, so that \begin{equation}\lb{9}( N_{p_e})^t\omega(p_e)+\omega(p_e)N_{p_e}=0, \end{equation} and inserting ~\rf{3} and~\rf{5} into~\rf{9} gives \begin{equation}\lb{10}N_{p_e}^{21}=-(N_{p_e}^{13})^*\omega_{\mu_e}^\flat, \qquad(N_{p_e}^{23})^*=N_{p_e}^{23}, \end{equation} so $N_{p_e}^{23}$ is symmetric, and in a certain sense, $N_{p_e}^{13}$ and $N_{p_e}^{21}$ are dual. Similarly, using~\rf{4} and~\rf{5} to write out what it means for $S_{p_e}$ and~$ N_{p_e}$ to commute, \begin{gather} \lb{14}dX_{H_{\mu_e}}(\bar p_e)N_{p_e}^{13}=-N_{p_e}^{13} \onm{coad_{\xi_e}},\\ \lb{15}N_{p_e}^{21}dX_{H_{\mu_e}}(\bar p_e)=-\onm{ad}_{\xi_e}N_{p_e}^{21},\\ \lb{16}\onm{ad}_{\xi_e}N_{p_e}^{23}=N_{p_e}^{23}\onm{coad}_{\xi_e}. \end{gather} These properties will yield the important \emph{normal form symmetry} of the drift system. Temporarily let $\onm{pr}_2$ be the projection onto the second factor of~$\onm{ker}dJ(p_e)=W_{\!\mathrm{red}}\oplus \mathfrak g_{\mu_e} p_e$. Then \begin{equation*} \onm{ker}\bigl(dJ(p_e)|(W_{\!\mathrm{red}}+\mathfrak bp_e)^{\omega\perp}\bigr) =\mathfrak g_{\mu_e} p_e\subseteq\onm{ker}\bigl(\onm{pr}_2N_{p_e}| (W_{\!\mathrm{red}}+\mathfrak bp_e)^{\omega\perp}\bigr), \end{equation*} so $N_{p_e}^{23}:\mathfrak g_{\mu_e}\times\mathfrak g_{\mu_e}\rightarrow\mathbb R$ is the unique bilinear from such that \begin{equation}\lb{17}\begin{diagram} \node{(W_{\!\mathrm{red}}+\mathfrak bp_e)^{\omega\perp}}\arrow{e,t}{dJ(p_e)} \arrow{s,l}{N_{p_e}} \node{\mathfrak g_{\mu_e}^*}\arrow{ssw,b}{(N_{p_e}^{23})^\flat}\\ \node{\onm{ker}dJ(p_e)}\arrow{s,l}{\onm{pr}_2}\\ \node{\mathfrak g_{\mu_e}} \end{diagram}\end{equation} Thus, $N^{23}$ may be calculated without calculating the Lagrangian complement $Z$ of~\rf{1}. However, $N_{p_e}^{23}$ \emph{does} depend on the particular choice or $W_{\!\mathrm{red}}$, even if $W_{\!\mathrm{red}}$ is an $S_{p_e}$-invariant complement of $\mathfrak g_{\mu_e} p_e$ in $\onm{ker}dJ(p_e)$. To see this dependence, choose a linear map $A:W_{\!\mathrm{red}}\rightarrow \mathfrak g_{\mu_e}$, which commutes with $S_{p_e}$, and then choose a new $W_{\!\mathrm{red}}$ where \begin{equation}\lb{12} W_{\!\mathrm{red}}^{\,\prime}=\bset{w+(Aw)p_e}{w\inW_{\!\mathrm{red}}}. \end{equation} Also, choose a map $B:\mathfrak b\rightarrow\mathfrak g_{\mu_e}$ which intertwines the adjoint action and then choose a new $\mathfrak b^\prime$ by \begin{equation}\lb{11} \mathfrak b^\prime=\bset{\eta+B\eta}{\eta\in\mathfrak b}. \end{equation} Given $\nu\in\mathfrak g_{\mu_e}^*$, the process of calculating $({N_{p_e}^{23}}^\prime)^\flat\nu$, given by the analog of~\rf{17} for the new choices, is \begin{enumerate} \item pick $z^\prime\in (W_{\!\mathrm{red}}^{\,\prime}+\mathfrak b^\prime p_e)^{\omega\perp}$ such that $dJ(p_e)z^\prime=\nu$; \item calculate $N_{p_e}z^\prime$; \item project $N_{p_e}z^\prime$ to $\mathfrak g_{\mu_e} p_e\equiv\mathfrak g_{\mu_e}$ using the splitting of $\onm{ker}dJ(p_e)$ defined by $W_{\!\mathrm{red}}^{\,\prime}$, with result $({N_{p_e}^{23}}^\prime)^\flat\nu$. \end{enumerate} It is straightforward from~\rf{3},~\rf{12}, and~\rf{11}, that \begin{equation*} (W_{\!\mathrm{red}}^{\,\prime}+\mathfrak b^\prime p_e)^{\omega\perp} =\bset{\omega_{\mu_e}^\sharp A^*\mu\oplus\xi\oplus\mu\oplus\breve\omega_{\mu_e} ^\sharp B^*\mu}{\xi\in\mathfrak g_{\mu_e},\mu\in\mathfrak g_{\mu_e}^*}, \end{equation*} so given $\nu\in\mathfrak g_{\mu_e}^*$, let \begin{equation*} z^\prime=\omega_{\mu_e}^\sharp A^*\nu\oplus0\oplus\nu\oplus\breve \omega_{\mu_e}^\sharp B^*\nu.\end{equation*} Then \begin{equation*}\begin{split} N_{p_e} z^\prime&= N_{p_e}^{13}\nu\oplus(N_{p_e}^{21}\omega_{\mu_e}^\sharp A^*\nu +(N_{p_e}^{23})^\flat\nu)\oplus0\oplus0\\ &=N_{p_e}^{13}\nu\oplus AN_{p_e}^{13}\nu\oplus0\oplus0\\ &\qquad\mbox{}+0\oplus(N_{p_e}^{21}\omega_{\mu_e}^\sharp A^*\nu -AN_{p_e}^{13}\nu+(N_{p_e}^{23})^\flat\nu)\oplus0\oplus0, \end{split}\end{equation*} which by~\rf{12} is the appropriate decomposition of $N_{p_e}z^\prime$. Thus, using~\rf{10}, \begin{equation}\lb{13}\begin{split} ({N_{p_e}^{23}}^\prime)^\flat\nu&= N_{p_e}^{21}\omega_{\mu_e}^\sharp A^*\nu -AN_{p_e}^{13}\nu+(N_{p_e}^{23})^\flat\nu\\ &=-\bigl((N_{p_e}^{13})^*A^*+AN_{p_e}^{13}\bigr)\nu +(N_{p_e}^{23})^\flat\nu. \end{split}\end{equation} This freedom to adjust $N_{p_e}^{23}$ by adjusting $W_{\!\mathrm{red}}$ will give the important \emph{gauge freedom} of the drift system. This gauge freedom can be encoded as a group, as follows. Temporarily set $\mathbb E_{p_e}=T_{\bar p_e}P_{\mu_e}\times\frak g_{\mu_e}\times\frak g_{\mu_e}^*\times T_{\mu_e} G_{\mu_e}$, and fix one particular splitting of type~\rf{1}, giving a symplectomorphism, say $\phi_0:T_{p_e}P\rightarrow\mathbb E_{p_e}$. Given $A:W_{\!\mathrm{red}}\rightarrow\frak g_{\mu_e}$ and $B:\frak b\rightarrow\frak g_{\mu_e}$ as above, there is another symplectomorphism $\phi_{A,B}:T_{p_e}P\rightarrow\mathbb E_{p_e}$, and hence a unique $\Delta_{A,B}$ in the symplectic group $Sp(\mathbb E_{p_e})$ such that \begin{equation*}\begin{diagram} \node[2]{T_{p_e}P}\arrow{sw,l}{\phi_0}\arrow{se,t}{\phi_{A,B}}\\ \node{\mathbb E_{p_e}}\arrow[2]{e,b}{\Delta_{A,B}}\node[2]{\mathbb E_{p_e}} \end{diagram}\end{equation*} Conversely, $A$ is determined by $\Delta_{A,B}$, since if $x\in T_{\bar p_e}P_{\mu_e}$, $\xi\in\frak g_{\mu_e}$, and $w\inW_{\!\mathrm{red}}$ is such that $\phi_0(w)=x$, then \begin{multline*} \Delta_{A,b}(x)=\phi_{A,B}(w) =\phi_{A,B}\bigl(w+(Aw)p_e-(Aw)p_e\bigr)\\= T\pi_{\mu_e}\bigl(w+(Aw)p_e\bigr)\oplus(-Aw) =x\oplus(-Aw). \end{multline*} Similarly $B$ is determined by $\Delta_{A,B}$, so the pairs $(A,B)$ have a natural group structure given by the injection $(A,B)\mapsto\phi_{A,B}$, and a simple calculation gives \begin{equation*} \phi_{A+A^\prime,B+B^\prime}=\phi_{A,B}\circ\phi_{A^\prime,B^\prime}. \end{equation*} Thus, at this level, the \emph{gauge group} is the additive Abelian group of pairs $(A,B)$ such that $A$ commutes with $S_{p_e}$ and $B$ commutes with $\onm{ad}_{\xi_e}$. \section{The simplest resonant case} All the above has been established without any particular presumption on the way that the group and reduced spectra intersect. There are some apriori features of the spectra: the spectra are purely imaginary (or zero) and are invariant under change of sign. Since $p_e$ is presumed to be formally stable, zero cannot occur in the reduced spectrum, and hence not in the intersection of the spectra. Thus, the following \defemph{resonance condition} is the simplest possible case beyond an empty intersection: \begin{equation*}\parbox{.8\displaywidth}{ \em The intersection of the spectrum of $\onm{ad}_{\xi_e}$ and the spectrum of $dX_{H_{\mu_e}}(\bar p_e)$ is $\{\pm i\lambda_{p_e}\}$, $\lambda_{p_e}>0$, and each of the eigenvalues $\pm i\lambda_{p_e}$ occur in each of these operators with multiplicity one.} \end{equation*} While many of the results below remain true or have analogues in a more relaxed environment, for simplicity, this resonance condition will be assumed for the remainder of this article. \subsection{The linearization} Given the resonance assumption, it is natural to consider certain spectral splittings of the factors $W_{\!\mathrm{red}}$ and~$\mathfrak g_{\mu_e}$ in~\rf{2}. In particular, let $W_{\!\mathrm{red}}^1$ be the generalized eigenspace of $\pm i\lambda_{p_e}$ for $dX_{H_{\mu_e}}(\bar p_e)$, and $W_{\!\mathrm{red}}^0$ be the sum of the other generalized eigenspaces, so that $W_{\!\mathrm{red}}^1$ is two dimensional, symplectic, and there is the symplectic splitting \begin{equation}\lb{101} T_{\bar p_e}P_{\mu_e}=W_{\!\mathrm{red}}^0\oplusW_{\!\mathrm{red}}^1. \end{equation} Similarly, let $\mathfrak g_{\mu_e}^1$ be the generalized eigenspace of $\pm i\lambda_{p_e}$ for the linear map $\onm{ad}_{\xi_e}$, and $\mathfrak g_{\mu_e}^0$ be the sum of the other generalized eigenspaces, so $\mathfrak g_{\mu_e}^1$ is two dimensional, and \begin{equation}\lb{102} \mathfrak g_{\mu_e}=\mathfrak g_{\mu_e}^0\oplus\mathfrak g_{\mu_e}^1. \end{equation} Of course, there is also then the dual splitting $\mathfrak g_{\mu_e}^*=\mathfrak g_{\mu_e}^{0*}\oplus\mathfrak g_{\mu_e}^{1*}$. So we have a refinement of the splitting~\rf{2} into ``nonresonant'' (with superscript~0) and ``resonant'' (superscript~1, dimension~2) parts: \begin{align} T_{p_e}P&=W_{\!\mathrm{red}}^0\oplusW_{\!\mathrm{red}}^1\oplus\mathfrak g_{\mu_e}\oplus \mathfrak g_{\mu_e}^*\oplus T_{\mu_e}(G{\mu_e})\lb{27}\\ &=W_{\!\mathrm{red}}^0\oplusW_{\!\mathrm{red}}^1\oplus\mathfrak g_{\mu_e}^0\oplus\mathfrak g_{\mu_e}^1\oplus \mathfrak g_{\mu_e}^{0*}\oplus\mathfrak g_{\mu_e}^{1*}\oplus T_{\mu_e}(G{\mu_e}).\nonumber \end{align} As one might expect, $N_{p_e}^{13}$ and $N_{p_e}^{21}$ localize to the resonant parts: \begin{alignat}{2} &W_{\!\mathrm{red}}^0\subseteq\onm{ker}N_{p_e}^{21},& \qquad &\onm{Image}N_{p_e}^{21}\subseteq\mathfrak g_{\mu_e}^1,\lb{19}\\ &\mathfrak g_{\mu_e}^0\subseteq\onm{ker}N_{p_e}^{13},&\qquad&\onm{Image} N_{p_e}^{13}\subseteqW_{\!\mathrm{red}}^1.\lb{20} \end{alignat} Indeed, to show the first of~\rf{19}, it suffices to find a subspace, say ${W_{\!\mathrm{red}}^0}^{\!\!\!\prime}\;$ of $\onm{ker}dJ(p_e)$ such that \begin{equation}\lb{21} T_{p_e}\pi_{\mu_e}{W_{\!\mathrm{red}}^0}^{\!\!\!\prime}\;=W_{\!\mathrm{red}}^0\quad\mbox{and}\quad N_{p_e}|{W_{\!\mathrm{red}}^0}^{\!\!\!\prime}\; = 0. \end{equation} The result then follows by using ${W_{\!\mathrm{red}}^0}^{\!\!\!\prime}\;$ to reverse the vertical arrow of~\rf{24}. The first of~\rf{21} can be assured by setting ${W_{\!\mathrm{red}}^0}^{\!\!\!\prime}\;$ to be the sum of the of nonresonant (i.e. not $\pm i\lambda_{p_e}$) generalized eigenspaces of $dX_{H_{\xi_e}}(p_e)$, and the second of~\rf{21} follows since ${W_{\!\mathrm{red}}^0}^{\!\!\!\prime}\;$ is $dX_{H_{\xi_e}}(p_e)$-invariant and that $dX_{H_{\xi_e}}(p_e)|{W_{\!\mathrm{red}}^0}^{\!\!\!\prime}\;$ is semisimple. For the second of~\rf{19}, $N_{p_e}$ maps the $\pm i\lambda_{p_e}$ generalized eigenspace of $dX_{H_{\xi_e}}(p_e)$ to itself, since $dX_{H_{\xi_e}}(p_e)$ does that. Thus the image of $N_{p_e}$ is contained in the intersection of that generalized eigenspace with $\mathfrak g_{\mu_e} p_e$, which is exactly $\mathfrak g_{\mu_e}^1 p_e$. From~\rf{19} one gets~\rf{20} by using the duality~\rf{10} between $N_{p_e}^{21}$ and~$N_{p_e}^{13}$. With respect to~\rf{27}, and in view of~\rf{5},~\rf{19} and~\rf{20}, the linear map $N_{p_e}$ has the form \begin{equation*} N_{p_e}=\left[\begin{array}{ccccc} 0&0&0&0&0\\ 0&0&0&N_{p_e}^{131}&0\\ 0&N_{p_e}^{211}&0&N_{p_e}^{23}&0\\ 0&0&0&0&0\\ 0&0&0&0&0\end{array}\right] \end{equation*} where $N_{p_e}^{211}\equiv N_{p_e}^{21}|W_{\!\mathrm{red}}^1$ and where $N_{p_e}^{131}$ is $N_{p_e}^{13}$ regarded as a map into $W_{\!\mathrm{red}}^1$. Let the quadratic Hamiltonian for the infinitesimally linear map $ N_{p_e}$ be $H_{\mathrm{nil}}$. With respect to~\rf{27}, the array for $H_{\mathrm{nil}}$ is \begin{multline*} \left[\begin{array}{ccccc} 0&0&0&0&0\\ 0&0&0&N_{p_e}^{131}&0\\ 0&N_{p_e}^{211}&0&N_{p_e}^{23}&0\\ 0&0&0&0&0\\ 0&0&0&0&0\end{array}\right]^* \left[\begin{array}{ccccc} \omega_{\mu_e}^0(\bar p_e)&0&0&0&0\\ 0&\omega_{\mu_e}^1(\bar p_e) &0&0&0\\ 0&0&0&\Id&0\\ 0&0&-\Id&0&0\\ 0&0&0&0&\breve\omega_{\mu_e}(p_e) \end{array}\right]\\ =\left[\begin{array}{ccccc} 0&0&0&0&0\\ 0&0&0&(N_{p_e}^{211})^*&0\\ 0&0&0&0&0\\ 0&(N_{p_e}^{131})^*\omega_{\mu_e}^1(\bar p_e)&0&N_{p_e}^{23}&0\\ 0&0&0&0&0\end{array}\right],\qquad\qquad \end{multline*} where $\omega_{\mu_e}^i(\bar p_e)$ denotes the symplectic form on $W_{\!\mathrm{red}}^i$, $i=1,2$. Using~\rf{16}, and the fact that the projections from the splitting $\mathfrak g_{\mu_e}^*=\mathfrak g_{\mu_e}^{0*}\oplus\mathfrak g_{\mu_e}^{1*}$ are certain polynomials in $\onm{coad}_{\xi_e}$, the bilinear form $N_{p_e}^{23}$ block diagonalizes over that splitting. Denote the blocks by $N_{p_e}^{230}$ and $N_{p_e}^{231}$. Then, using the variables \begin{equation*} (x,\mu)=(x,\mu^0,\mu^1)\inW_{\!\mathrm{red}}^1\times\mathfrak g_{\mu_e}^*=W_{\!\mathrm{red}}^1\times\mathfrak g_{\mu_e}^{0*}\times\mathfrak g_{\mu_e}^{1*}, \end{equation*} we have \begin{equation}\lb{33}\begin{split} H_{\mathrm{nil}}&=\frac12N_{p_e}^{23}(\mu,\mu)+ \langle\mu,N_{p_e}^{211}x\rangle\\ &=\frac12N_{p_e}^{230}(\mu^0,\mu^0)+\frac12N_{p_e}^{231}(\mu^1,\mu^1) +\langle\mu^1,N_{p_e}^{211}x\rangle. \end{split}\end{equation} I now construct the normal form symmetry for $H_{\mathrm{nil}}$. Let $T_{\xi_e}$ be the closure of $\exp(\mathbb R\xi_e)$, so $T_{\xi_e}$ is Abelian, is generically a maximal torus of $G_{\mu_e}$, and $T_{\xi_e}$ naturally acts on both $\mathfrak g_{\mu_e}$ and $\mathfrak g_{\mu_e}^*$, with invariant subspaces $\mathfrak g_{\mu_e}^i$ and $\mathfrak g_{\mu_e}^{i*}$, $i=1,2$. By~\rf{16}, $N_{p_e}^{23}$ is $T_{\xi_e}$-invariant, and so~\rf{33} is invariant under $T_{\xi_e}$ restricted to the $\mu^0$ variables alone. After this, it is the last two terms of~\rf{33} that will dictate its further symmetries. Now $S^1$ also acts\footnote{Define $\theta^\wedge=\theta\left(\begin{array}{cc}0&-1\\1&0\end{array}\right)$ so that $S^1=SO(2)$ and $\exp$ is the usual matrix exponential.} on $\mathfrak g_{\mu_e}^1$ by \begin{equation*} \exp(\theta^\wedge)\eta\equiv\exp\left(\frac\theta{\lambda_{p_e}}{\xi_e} \right)\eta,\quad\theta\in\mathbb R, \end{equation*} which is merely a renormalization of the action of $T_{\xi_e}$ on $\mathfrak g_{\mu_e}^1$, and by duality there is a corresponding action of $S^1$ on $\mathfrak g_{\mu_e}^*$. Also there is the following symplectic action of $S^1$ on $W_{\!\mathrm{red}}^1$: \begin{equation}\lb{107} \exp(\theta^\wedge) x\equiv\exp\left(\frac\theta{\lambda_{p_e}} dX_{H_{\mu_e}}(\bar p_e)\right)x,\quad\theta\in\mathbb R. \end{equation} By~\rf{14} and~\rf{15}, $N_{p_e}^{131}$ and $N_{p_e}^{211}$ reverse-intertwine the two $S^1$ actions, so that \begin{equation*} \langle{\exp(\theta^\wedge)}^{-1}\mu^1, N_{p_e}^{211}\exp(\theta^\wedge) x\rangle =\langle{\exp(\theta^\wedge)}^{-1}\mu^1, {\exp(\theta^\wedge)}^{-1} N_{p_e}^{211}x\rangle=\langle\mu^1,x\rangle. \end{equation*} Consequently, the \defemph{normal form symmetry} will be $T_{\xi_e}\times S^1$ where the first factor acts on the $\mu^0$ variables via the coadjoint action and the second factor acts by \begin{equation*} a\cdot(x,\mu^1)\equiv(a^{-1}x,a\mu),\quad a=\exp(\theta^\wedge). \end{equation*} The normal form symmetry restricts the quadratic Hamiltonian~$H_{\mathrm{nil}}$, as follows. There is a basis of the linear space $W_{\!\mathrm{red}}^1=\{(x_1,x_2)\in\mathbb R^2\}$ so that the symplectic form is $dx^1\wedge dx^2$ and the $S^1$ action on $W_{\!\mathrm{red}}^1$ is either $\bigl(\exp(\theta^\wedge),x\bigr)\mapsto \exp(\theta^\wedge)x$ or $\bigl(\exp(\theta^\wedge),x\bigr)\mapsto \exp(-\theta^\wedge)x$ (i.e. counterclockwise or clockwise rotation). Using the $\onm{Ad}$-invariant metric of $\mathfrak g$, choose a orthonormal basis for the space $\mathfrak g_{\mu_e}^1$, so that $\mathfrak g_{\mu_e}^1=\{(\pi_1,\pi_2)\in\mathbb R^2\}$ and the $S^1$ action on $\mathfrak g_{\mu_e}^1$ is a rotation in the opposite sense as the $S^1$ action on $W_{\!\mathrm{red}}$. The map $N_{p_e}^{211}$ is highly restricted: since it reverse-intertwines these two reverse $S^1$ actions, there is a $\kappa$ such that \begin{equation} N_{p_e}^{211}=\kappa\left[\begin{array}{cc}1&0\\0&1\end{array}\right] \qquad\mbox{or}\qquad N_{p_e}^{211}=\kappa\left[\begin{array}{cc}0&-1\\1&0\end{array}\right]. \lb{40} \end{equation} Equation~\rf{40} defines $\kappa$, so that the map $N^{211}$ has the effect of giving rise to this single ``coupling constant'' and one of the two possible choices in~\rf{40}. Without loss of generality, by the symplectic isomorphisms $(x_1,x_2)\mapsto(-x_1,-x_2)$ and $(x_1,x_2)\mapsto(-x_2,x_1)$ on $W_{\!\mathrm{red}}^1$ , $\kappa\ge0$ and the second choice in~\rf{40} may be discarded, respectively. Then, since $N^{23}_{p_e}$ is $S^1$-symmetric, a constant $I_1$ may be obtained from $N_{p_e}^{231}$ such that \begin{equation}\lb{100} H_{\mathrm{nil}}= \frac12N_{p_e}^{230}(\mu^0,\mu^0) +\frac12I_1({\pi_1}^2+{\pi_2}^2) +\kappa(\pi_1x_1+\pi_2x_2). \end{equation} According to the above conventions, $N^{211}$ reverse intertwines reverse-sense $S^1$ actions on the $(x_1,x_2)$ variables and the $(\pi_1,\pi_2)$ variables, and the second action has the same sense as the action of $\onm{ad}_{\xi_e}$. Consequently, {\em there are two distinct possibilities:} the $S^1$ action on the $(x_1,x_2)$ variables can be counterclockwise or clockwise. To avoid unnecessary signs that would otherwise appear later, I will redefine the $S^1$ actions in the following way: if the $S^1$ action on the $(x_1,x_2)$ variables is clockwise then reverse it, and both actions are then counterclockwise and the $S^1$ action on the $(\pi_1,\pi_2)$ variables is in the same sense as the action of $\onm{ad}_{\xi_e}$. If on the other hand the $S^1$ action on the $(x_1,x_2)$ variables is counterclockwise then the action of the $(\pi_1,\pi_2)$ variables is clockwise, and is to be reversed, so that it has the opposite sense to the action of $\onm{ad}_{\xi_e}$. Thus, by these conventions, both actions are counterclockwise, $N^{211}$ is intertwining, and the distinctness of the two cases appears as one of two possibilities: the $S^1$ action on the $(\pi_1,\pi_2)$ variables might be in the same sense as the action of $\onm{ad}_{\xi_e}$ (I call this the ``$+$'' case, and it occurs when the original action on the $(x_1,x_2)$ variables is clockwise) or the opposite sense as the action of $\onm{ad}_{\xi_e}$ (the ``$-$'' case, occurring when the original action on the $(x_1,x_2)$ variables is counterclockwise). Although it may appear that same-sense vs. opposite-sense with respect to $\onm{ad_{\xi_e}}$ holds some significance, similar to, for example, the difference between a $1:1$ and $1:-1$ resonance as in~\ctt{KummerM-1976.1}{KummerM-1978.1}, this is not so: changing the the symplectic form $dx_1\wedge dx_2$ to $dx_2\wedge dx_1$ will exchange my ``$+$'' and ``$-$'' cases. These two cases are qualitatively identical and the subsequent theory is ambidextrous with respect to them. Now I will determine the effect on~\rf{100} of choosing various subspaces $W_{\!\mathrm{red}}$, or equivalently, various operators~$A$ commuting with $S_{p_e}$ in~\rf{12}. With respect to the decompositions~\rf{101} and~\rf{102} let $A$ have the block form \begin{equation*} A=\left[\begin{array}{cc}A^{11}&A^{12}\\A^{21}&A^{22}\end{array}\right]. \end{equation*} To determine which $A$ commute with $S_{p_e}$, note that $S_{p_e}$ has a block diagonal form on $W_{\!\mathrm{red}}^0\oplusW_{\!\mathrm{red}}^1\oplus\mathfrak g_{\mu_e}^0\oplus\mathfrak g_{\mu_e}^1$ since~\rf{101} and~\rf{102} are generalized eigenspace decompositions for $S_{p_e}$; let the blocks be $S^{110}$, $S^{111}$, $S^{220}$, and $S^{221}$, so $AS=SA$ becomes \begin{equation}\lb{105} \left[\begin{array}{cc}A^{11}&A^{12}\\A^{21}&A^{22}\end{array}\right] \left[\begin{array}{cc}S^{110}&0\\0&S^{111}\end{array}\right]= \left[\begin{array}{cc}S^{220}&0\\0&S^{221}\end{array}\right] \left[\begin{array}{cc}A^{11}&A^{12}\\A^{21}&A^{22}\end{array}\right]. \end{equation} When multiplied, the top left blocks of~\rf{105} give $A^{11}S^{110}=S^{220}A^{11}$, which is possible only if $A^{11}$ is zero, since $S^{110}$ and $S^{220}$ have no common spectrum, and similarly $A^{12}=0$ and $A^{21}=0$. However, $S^{111}$ and $S^{221}$ both have spectrum $\{\pm\lambda_e\}$, so one cannot conclude $A^{22}=0$, but only that $A^{22}S^{111}=S^{221}A^{22}$. Since $S^{111}$ is exactly the linearization~$dX_{H_{\mu_e}}(\bar p_e)$ restricted to $W_{\!\mathrm{red}}^1$, which by~\rf{107} defined the $S^1$ action on $W_{\!\mathrm{red}}^1$, which is counterclockwise or clockwise, we have \begin{equation*} S^{111}=\pm\sqrt{\lambda_e}\left[\begin{array}{cc}0&1\\-1&0 \end{array}\right]. \end{equation*} Since $S^{221}$ is $-\onm{ad}_{\xi_e}$ on $\mathfrak g_{\mu_e}^1$, and $\onm{ad}_{\xi_e}$ generates the $S^1$ action of opposite sense to that generated by $S^{111}$, we have \begin{equation*} S^{221}=\pm\sqrt{\lambda_e}\left[\begin{array}{cc}0&1\\-1&0 \end{array}\right]. \end{equation*} By letting \begin{equation*} A^{22}=\left[\begin{array}{cc}a^{11}&a^{12}\\a^{21}&a^{22} \end{array}\right], \end{equation*} $A^{22}S^{111}=S^{221}A^{22}$ becomes $a^{12}=-a^{21}$ and $a^{11}=a^{22}$. By~\rf{40} and the duality~\rf{10}, \begin{equation*} N_{p_e}^{13}=\left[\begin{array}{cc}0&0\\0& \left[\begin{array}{cc}0&\kappa\\-\kappa&0\end{array}\right] \end{array}\right], \end{equation*} so that \begin{equation*} (N_{p_e}^{13})^*A^*+AN_{p_e}^{13}= \left[\begin{array}{cc}0&0\\0& -\kappa\left[ \begin{array}{cc}2a^{12}&0\\0&2a^{12} \end{array}\right] \end{array}\right], \end{equation*} so by~\rf{13}, the Hamiltonian generated by the new choice of $W_{\!\mathrm{red}}$ has the same form as~\rf{100} but with $I_1$ replaced by $I_1+2\kappa a^{12}$. Thus, {\em the entire effect of the freedom to choose~$W_{\!\mathrm{red}}$ is exactly that the constant $I_1$ may be arbitrarily manipulated, as long as $\kappa\ne0$. The effective action of the gauge group as it acts on the normal form is isomorphic to the additive group of real numbers.} \subsection{The normal form} The splitting of $T_{p_e}P$ and subsequent analysis of the linearization $dX_{H_{\xi_e}}(p_e)$ amounts to the construction of a normal form for linearizations of relative equilibria. When equivariantly moved through $G_{\mu_e} p_e$, this normal form can be combined with the equivariant isotropic embedding theorem~(\ct{MarsdenJE-1981.1},\ct{WeinsteinA-1997.1}), to provide a $G_{\mu_e}$-equivariant normal form for the entire Hamiltonian system near $G_{\mu_e} p_e$. I now construct this normal form, closely following~\ct{PatrickGW-1995.1}. Let $\tilde P$ be the symplectic manifold \begin{equation*}\tilde P\equiv W_{\!\mathrm{red}}^0\timesW_{\!\mathrm{red}}^1\times T^*G_{\mu_e} \times G{\mu_e},\end{equation*} and define \begin{equation*}\tilde p_e\equiv 0\oplus0\oplus0_\Id\oplus{\mu_e}. \end{equation*} Let the group $G_{\mu_e}$ act on $\tilde P$ by left translation on the third factor. At the zero section of $T^*G_{\mu_e}$, there is the splitting of the tangent space of $T^*G_{\mu_e}$ into vertical and horizontal parts, and hence \begin{equation}\lb{26} T_{\tilde p_e}\tilde P=W_{\!\mathrm{red}}^0\oplusW_{\!\mathrm{red}}^1\oplus\mathfrak g_{\mu_e}\oplus \mathfrak g_{\mu_e}^*\oplus T_{\mu_e}(G{\mu_e}).\end{equation} Together,~\rf{27} and~\rf{26} yield a linear symplectomorphism between $T_{p_e}P$ and $T_{\tilde p_e}\tilde P$, which extends by $G_{\mu_e}$-equivariance to a vector bundle isomorphism, say $\Lambda^\prime:TP|(G_{\mu_e} p_e)\rightarrow T\tilde P|(G_{\mu_e} \tilde p_e)$, which is symplectic on each fiber. The equivariant isotropic embedding theorem then extends $\Lambda^\prime$ to an equivariant symplectomorphism $\Lambda$ from a neighborhood of $G_{\mu_e} p_e$ to a neighborhood of $G_{\mu_e}\tilde p$. The relationship between $\Lambda^\prime$ and $\Lambda$ is $\Lambda^\prime=T\Lambda$ on the domain of $\Lambda^\prime$. The linearization $dX_{H_{\xi_e}}(p_e)$ is in a sense a first order approximation of $X_{H_{\xi_e}}$ at $p_e$, and the quadratic Hamiltonian associated to $dX_{H_{\xi_e}}(p_e)$ is a second order approximation to $H_{\xi_e}$ at $p_e$. Together with the map $\Lambda$, this yields a second order approximation to the Hamiltonian~$\tilde H$ in a neighborhood of $G_{\mu_e} \tilde p_e$. Indeed, let $\tilde J^\xi$ be the momentum mapping for the left action of right translations $(g,h)\mapsto hg^{-1}$, so that \begin{equation*} \tilde J^\xi(\alpha_g)\equiv-\langle\alpha_g,X_\xi(g)\rangle, \end{equation*} where $X_\xi$ is the left invariant vector field generated by $\xi$. Then \begin{equation}\lb{34}\tilde H=H_{\mathrm{nil}}-(\tilde J^{\xi_e}-H_{\mathrm{red}}^1) +H_{\mathrm{red}}^0+H_{\mathrm{rsd}}+H_{\mathrm{rmd}}, \end{equation} where \begin{enumerate} \item $H_{\mathrm{nil}}:W_{\!\mathrm{red}}^1\times T^* G_{\mu_e}\rightarrow\mathbb R$ is the extension of~\rf{33} to $T^* G_{\mu_e}$ by left invariance, \item $H_{\mathrm{red}}^i:W_{\!\mathrm{red}}^i\rightarrow\mathbb R$ is $d^2 H_{\mu_e}(\bar p_e)|W_{\!\mathrm{red}}^i$, $i=1,2$, \item $H_{\mathrm{rsd}}:G\mu_e\rightarrow\mathbb R$ by $H_{\mathrm{rsd}}(g\mu_e)\equiv-\langle g\mu_e, \xi_e\rangle$, and \item $dH_{\mathrm{rmd}}=0$ and $d^2H_{\mathrm{rmd}}=0$ on $G_{\mu_e} \tilde p_e$. \end{enumerate} Items (1)--(3) are definitions, while the content is in item~(4), which follows since the two Hamiltonians \begin{equation*} \tilde H, \qquad H_{\mathrm{nil}}-(\tilde J^{\xi_e}-H_{\mathrm{red}}^1)+H_{\mathrm{red}}^0+H_{\mathrm{rsd}} \end{equation*} both have relative equilibria at $\tilde p_e$ and have been manipulated to have matched linearizations (and consequently matched Hessians) at each point of $G_{\mu_e}\tilde p_e$. The point is that a second order approximation to $\tilde H$ may be obtained by dropping the remainder term $H_{\mathrm{rmd}}$ of~\rf{34}. Now, consider $\tilde H$ as defined by~\rf{34}. The second term (the one grouped within brackets) generates the one parameter group of symplectomorphisms generated by $\xi_e$ through the action~\rf{38}. Therefore, the first term $H_{\mathrm{nil}}$ and the second term Poisson commute. Obviously, then, the first four terms of~\rf{34} pairwise Poisson commute. The effect of the term $-\tilde J^{\xi_e}$ is merely to generate the flow of the right invariant vector field generated by ${\xi_e}$, and this corresponds to the ``fast'' evolution of the relative equilibrium $\tilde p_e$, while the flows of the terms $H_{\mathrm{red}}^0$ and $H_{\mathrm{rsd}}$ cannot cause evolution along $G_{\mu_e}\tilde p_e$. So the slow evolution near $G_{\mu_e}\tilde p_e$ is to first order dictated by the flow of $H_{\mathrm{nil}}$. As for the symmetries of $\tilde H$, since the Hamiltonian $H$ is $G$ invariant and $\Lambda^\prime$ is $G_{\mu_e}$-equivariant, the Hamiltonian~\rf{34} is $G_{\mu_e}$-invariant, including the remainder term. Also, when $H_{\mathrm{nil}}$ is extended to be a left invariant function of $W_{\!\mathrm{red}}^1\times T^*G_{\mu_e}$, invariance under the normal form action given by~\rf{11} becomes right invariance, i.e., for the $S^1$ part, invariance under the action \begin{equation}\lb{38} a\cdot(x,\alpha_g)=\bigl(a^{-1}x,(TR_a)^*\alpha_g\bigr), \end{equation} where $R$ denotes right translation of $G_{\mu_e}$, and the functions $H_{\mathrm{red}}^0$ and $H_{\mathrm{rsd}}$ are also invariant under~\rf{38}. However this normal form symmetry does not in general extend to the remainder term $H_{\mathrm{rmd}}$. The action~\rf{38} is counterclockwise or clockwise diagonal rotation on the variables $(\pi_1,\pi_2)$ and $(x_1,x_2)$ in a way that is a sense-preserving renormalization of the coadjoint action of $\exp(\mathbb R\xi_e)$ on $(\pi_1,\pi_2)$. Since~\rf{38} is the action of the Abelian $S^1$, this action can be reversed and thus assumed to be counterclockwise; however, if this is done then the resulting action may acquire the {\em opposite} sense to the coadjoint action of $\exp(\mathbb R\xi_e)$ on $(\pi_1,\pi_2)$. In concept the map $\Lambda$ is like a coordinate system near $G_{\mu_e} p_e$; it is only that its range is a manifold rather than an open subset of Euclidean space. Through its dependence on the splitting of $T_{p_e}P$, $\Lambda$ depends on the choice of $W_{\!\mathrm{red}}$, so the effect of varying $W_{\!\mathrm{red}}$ like varying a coordinate system near $G_{\mu_e} p_e$. Thus, the freedom to vary $W_{\!\mathrm{red}}$ is like a gauge freedom. I will call a particular choice of $W_{\!\mathrm{red}}$ \defemph{ a choice of gauge}, and say that, in view of Section~(3.1), the gauge group of $p_e$ is $(\mathbb R,+)$. As shown in Section~(3.1), varying $W_{\!\mathrm{red}}$ itself is a linear process, the effect of which on $H_{\mathrm{nil}}$ is easily found and described. Yet, through the isotropic embedding theorem, $\Lambda$ depends on $W_{\!\mathrm{red}}$ in a nontrivial way. This ability to cause very nontrivial ``coordinate changes'' while easily calculating the concomitant effect on $H_{\mathrm{nil}}$ will enter powerfully in the subsequent analysis. Particularly, one can imagine choosing a gauge that ``simplifies'' $H_{\mathrm{nil}}$, making the analysis of $H_{\mathrm{nil}}$ tractable and yielding information about the flow of the original Hamiltonian system near $G_{\mu_e} p_e$. On the other hand, a simple $H_{\mathrm{nil}}$ will in general be nongeneric, so that higher order terms in $H_{\mathrm{rmd}}$ might destroy whatever delicate nongeneric structures arise from $H_{\mathrm{nil}}$. In that case, one can focus on the identification and perturbation of these nongeneric structures, which will occupy smaller regions of phase space as ones attention is restricted more nearly to $G_{\mu_e} \bar p_e$. {\em Summary: The phase space $P_{\mathrm{drift}}$ on which some Hamiltonian $H_{\mathrm{drift}}$ approximates to first order the evolution of the system, when that system is perturbed from $p_e$, is \begin{gather*} P_{\mathrm{drift}}=\mathbb R^2\times T^*G_{\mu_e},\\ H_{\mathrm{drift}}(x,\alpha_g)=\langle g^{-1}\alpha_g,\xi_e\rangle+ \frac12N_{p_e}^{23}(\alpha_g,\alpha_g)+ \langle g^{-1}\alpha_g, N_{p_e}^{211}x\rangle. \end{gather*} The equations of motion for this system\footnote{\lb{52} For a Hamiltonian system on the cotangent bundle of a Lie group $G$ of the form $\alpha_g\mapsto 1/2h(\alpha_g,\alpha_g)+\langle\alpha_g,X\rangle$ where $h$ is a left invariant metric on $G$ and $X$ is a left invariant vector field on $G$, one has the equations of motion \begin{equation*} \Omega\equiv h^\sharp\pi+X(e), \quad\pi\equiv g^{-1}\alpha_g,\quad g^{-1}\frac{dg}{dt}=\Omega,\quad \frac{d\pi}{dt}=-\onm{coad}_\Omega\pi. \end{equation*}} are \begin{equation*} g^{-1}\frac{dg}{dt}=\Omega,\quad\frac{d\pi}{dt}=\onm{coad}_\Omega\pi, \quad\frac{dx}{dt}=N^{131}_{p_e}\pi. \end{equation*} where \begin{equation*} \Omega\equiv\xi_e+(N^{23})^\flat\pi+N^{211}_{p_e}x. \end{equation*} There is a splitting $\mathfrak g_{\mu_e}=\{(\pi_1,\pi_2)\in \mathbb R^2\} \oplus\{\mu^0\}$ such that $H_{\mathrm{drift}}$ becomes, after Poisson reduction by left translation of $G_{\mu_e}$, \begin{equation}\lb{130} H_{\mathrm{drift}}= \langle\mu^0,\xi_e\rangle+\frac12N_{p_e}^{230}(\mu^0,\mu^0) +\frac12I_1({\pi_1}^2+{\pi_2}^2) +\kappa(\pi_1x_1+\pi_2x_2). \end{equation} Here $N^{230}$ is a constant quadratic form, and $I_1$ and $\kappa$ are constants, and all of these are calculable from the nilpotent part of the linearization of the relative equilibrium. The Hamiltonian~\rf{130} is invariant under counterclockwise diagonal rotation on the variables $(\pi_1,\pi_2)$ and $(x_1,x_2)$ and the coadjoint action of $T_{\xi_e}$ on the variables $\mu^0$. The gauge group of the approximation is $(\mathbb R,+)$, and the effect of the gauge freedom is to exactly to undetermine $I_1$.} \section{The mechanical case} Here I consider the special case $G=SO(3)\times (S^1)^n$. Standard identifications give $\mathfrak g=\mathbb R^3\times\mathbb R^n=\mathfrak g^*$. In order that the momentum be nongeneric, assume that the $SO(3)$ part of the momentum is zero. Reorient the system (i.e. left translate it) so that $\mathfrak g_{\mu_e}^1=\{(\pi_1,\pi_2,0)\}\subseteq so(3)$. Since the action of $\onm{ad}_{\xi_e}$ fixes $\mathfrak g_{\mu_e}^1$, the $SO(3)$ part of $\xi_e$ is either parallel or antiparallel to to $\mathbf k\in so(3)$, where $\mathbf k=(0,0,1)$. I drop the first term of~\rf{130}, since the background motion of the relative equilibrium $p_e$ itself will not be of interest, after which~\rf{130} becomes \begin{equation*}\begin{split} H_{\mathrm{drift}}=\frac12I_1\bigl({\pi_1}^2+{\pi_2}^2\bigr)+\frac12I_2{\pi_3}^2 &+\kappa(\pi_1x_1+\pi_2x_2)\\ &\quad+\pi_3 \sum_{i=1}^na_ip_i+\sum_{i,j=1}^na_{ij}p_ip_j, \end{split}\end{equation*} where the variable labeling is \begin{equation*}(\pi_1,\pi_2,\pi_3,p_1,\ldots,p_n)\in\mathfrak g^*= \mathbb R^3\times\mathbb R^n,\end{equation*} and $a_i$ and $a_{ij}$ are constants. The whole of $(S^1)^n$ is ignorable; an Abelian reduction yields the phase space and Hamiltonian (which by abuse of notation will have the same names) \begin{gather} P_{\mathrm{drift}}=\mathbb R^2\times T^*SO(3)=\bigl\{\bigl((x_1,x_2)\in\mathbb R^2, A\in SO(3),(\pi_1,\pi_2,\pi_3)\in\mathbb R^3\bigr)\bigr\} \notag\\ H_{\mathrm{drift}}=\frac12I_1({\pi_1}^2+{\pi_2}^2)+\frac12I_2{\pi_3}^2+ \kappa(\pi_1x_1+\pi_2x_2)+a\pi_3,\lb{41} \end{gather} where $a$ is constant. The normal form symmetry is counterclockwise diagonal action of $S^1$ on the variables $(x_1,x_2), (\pi_1,\pi_2)$, and the momentum mapping for the normal form symmetry is \begin{equation*}J^{\mbox{\scriptsize nf}}= -\frac12({x_1}^2+{x_2}^2)-\pi_3 \end{equation*} An easy verification (see the footnote on page~\pageref{52}) gives the equations of motion for~\rf{41} as \begin{equation} \frac d{dt}\left[\begin{array}{c}\pi_1\\\pi_2\\\pi_3\\x_1\\x_2 \end{array}\right]= \left[\begin{array}{c} I\pi_2\pi_3-\kappa x_2\pi_3+a\pi_2\\ -I\pi_1\pi_3+\kappa x_1\pi_3-a\pi_1\\ \kappa(\pi_1x_2-\pi_2x_1)\\ \kappa\pi_2\\-\kappa\pi_1\end{array}\right],\lb{54}\end{equation} and \begin{equation}\lb{55} A^{-1}\frac{dA}{dt}=\left[ \begin{array}{c}I_1\pi_1+\kappa x_1\\I_1\pi_2+\kappa x_2\\I_2\pi_3+a\end{array} \right]^\wedge, \end{equation} where $I\equiv I_2-I_1$ and as usual, for $v\in\mathbb R^3$, \begin{equation*}v ^\wedge\equiv\left[\begin{array}{ccc} \phantom{-}0\phantom{^3}&-v^3&\phantom{-}v^2\\ \phantom{-}v^3&\phantom{-}0\phantom{^2}&-v^1\\ -v^2&\phantom{-}v^1&\phantom{-}0\phantom{^2} \end{array}\right].\end{equation*} To obtain a first impression of~\rf{41} one can contrast it with the nonresonant case. On the one hand, the phase space $P_{\mathrm{drift}}$ of~\rf{41} has dimension~8; the symmetry group is $SO(3)\times S^1$. Hence, the Marsden-Weinstein reduced systems are integrable, since they have dimension $8-4-2=2$, and energy is a conserved quantity. We shall see that homoclinic connections appear in the reduced spaces of~\rf{41}, which are spheres in general, and that singular reduced spaces (pinched spheres) also occur for some momentum values. On the other hand, the {\em nonresonant\/} case (obtained from~\rf{41} by setting $\kappa=0$ and dropping the variables $x_1$ and $x_2$) has the same symmetry group but dimension~6 phase space; {\em it\/} is integrable by its symmetry group alone, without the use of energy. Dynamically, the nonresonant case is simpler, since every motion is a relative equilibrium and the compact symmetry group rules out homoclinic connections. \subsection{Full reduction} The Poisson manifold $\mathbb R^2\times\mathbb R^3=\{(x, \pi)\}$ may be reduced by the $S^1$ normal form symmetry by use of the Hopf variables (see, for example,~\ct{CushmanRHBatesLM-1997.1}, page~14; also page~407 has a summary of references about singular reduction): \begin{alignat*}{2} w_1&\equiv 2(x_1\pi_2-x_2\pi_1),&\qquad w_2&\equiv 2(x_1\pi_1+x_2\pi_2),\\ w_3&\equiv({x_1}^2+{x_2}^2)-({\pi_1}^2+{\pi_2}^2),&\qquad w_4&\equiv({x_1}^2+{x_2}^2)+({\pi_1}^2+{\pi_2}^2). \end{alignat*} This map is a quotient map for the $S^1$ normal form symmetry and it has the semialgebraic image given by the subset of $\mathbb R^5=\{(w_1,w_2,w_3,w_4,\pi_3)\}$ satisfying \begin{equation*} {w_1}^2+{w_2}^2+{w_3}^2-{w_4}^2=0,\quad w_4\ge 0. \end{equation*} The symplectic reduced spaces are the level sets of the two Casimirs~$j_1$ and $j_2$ given by \begin{gather} \lb{70} j_1\equiv({\pi_1}^2+{\pi_2}^2+{\pi_3}^2)^{\frac12}=\left({\pi_3}^2 +\frac12(w_4-w_3)\right)^\frac12,\\ \lb{71}j_2\equiv J^{\mbox{\scriptsize nf}}= -\frac12\bigl({x_1}^2+{x_2}^2\bigr)-\pi_3=-\frac14(w_3+w_4)-\pi_3. \end{gather} Given $j_1$ and $j_2$,~\rf{70} and~\rf{71} can be used to eliminate $w_3$ and $w_4$, thereby obtaining the symplectic reduced spaces as the subsets of $\mathbb R^3=\{(w_1,w_2,\pi_3)\}$ given by \begin{gather} {w_1}^2+{w_2}^2=8({\pi_3}^2-{j_1}^2)(\pi_3+j_2)\lb{73}\\ w_4=({j_1}^2-2j_2)-({\pi_3}^2+2\pi_3)\ge0.\lb{7300} \end{gather} Various symplectic reduced spaces are obtained by fixing various values of $j_1$ and $j_2$ , but the values of $j_1$ and $j_2$ are not arbitrary, since $j_1\ge0$ and \begin{equation*} j_1=({\pi_1}^2+{\pi_2}^2+{\pi_3}^2)^{\frac12}\ge-\pi_3=j_2+\frac12({x_1}^2+{x_2}^2) \ge j_2. \end{equation*} Moreover, the singularities in the symplectic reduced spaces occur where $x_1=x_2=\pi_1=\pi_2=0$, and here $j_1=|\pi_3|=\pm j_2$. Equation~\rf{73} has solutions only over the intervals where the cubic on the right hand side is nonnegative, and moreover, by~\rf{71}, $\pi_3\ge-j_2$. Putting all this together gives the bifurcation diagram shown in~Figure~\rf{703}. As shown there, the symplectic reduced spaces are the surfaces of revolution of $\bigl(8({\pi_3}^2-{j_1}^2)(\pi_3+j_2)\bigr)^{\frac12}$ over the {\em finite\/} interval where the cubic is positive. Thus, these spaces are points, spheres, or, in the case that $j_2=-j_1$, they are topological spheres with one conical singularity. \begin{figure}[p]\lb{200} \input fig1.tex\vspace*{.25in} \caption{\label{703}\it\protect\footnotesize The bifurcation diagram for the symplectic reduced spaces. Clockwise from the top, the reduced spaces are points, spheres, pinched spheres, again spheres and finally points. The finite intervals where the cubic is positive is the thicker black line on the $\pi_3$ axis. One verifies that Inequality~\rf{7300} is respected by showing that the concave down quadratic $w_4=-({\pi_3}^2+2\pi_3)-{j_1}^2+2j_2$ is positive on those intervals by checking the it is positive at the various endpoints of the intervals.} \end{figure} In passing, I note that the symplectic volumes of the $-j_2=j_1>0$ reduced phase spaces are infinite, and of course the infinity is concentrated at the singularity. In fact, by regarding two functions $f(w_1,w_2)$ and $g(w_1,w_2)$ as functions on one of those reduced phase spaces, lifting those functions to the Poisson phase space $\{(x,\pi)\}=\mathbb R^2\times so(3)$, and then calculating the Poisson bracket, one verifies that \begin{equation*} \{f,g\}=-4({\pi_3}^2+2j_2\pi_3+{j_1}^2)\left( \frac{\partial f}{\partial w_2}\frac{\partial g}{\partial w_1}- \frac{\partial f}{\partial w_1}\frac{\partial g}{\partial w_2}\right). \end{equation*} Consequently, the symplectic form on the reduced spaces in the coordinates $(w_1,w_2)$ is \begin{equation*} \omega_{j_1,j_2}=\frac1{4({\pi_3}^2+2j_2\pi_3+{j_1}^2)}\,dw_1\wedge dw_2. \end{equation*} Switching to polar coordinates $w_1=r\cos\theta$ and $w_2=r\sin\theta$, the symplectic volume of a small circle of radius $\delta$ about the origin is \begin{equation*} A_\delta\equiv2\pi\int_0^\delta\frac{r}{4({\pi_3}^2+2j_2\pi_3+{j_1}^2)}\,dr =\frac\pi2\int_0^\delta\frac{r}{(\pi_3-j_1)^2}\,dr \end{equation*} and upon substituting \begin{equation*} r^2=8({\pi_3}^2-{j_1}^2)(\pi_3-j_1)=8(\pi_3+j_1)(\pi_3-j_1)^2 \end{equation*} that symplectic volume becomes \begin{equation*} A_\delta=4\pi\int_0^a\frac{\pi_3+j_1}{r}\,dr. \end{equation*} Since the numerator tends to $2j_1$ as $r$ becomes small, $A_\delta$ is infinite. The Hamiltonian~\rf{41} on the reduced spaces is, with the aid of the Hopf variables and~\rf{70}, \begin{equation*}\begin{split} H_{\mathrm{drift}}&=\frac14I_1(w_4-w_3)+\frac12I_2{\pi_3}^2+ \frac\kappa2 w_2+a\pi_3\\ &=\frac12 I{\pi_3}^2+\frac\kappa2w_2+a\pi_3+\frac12I_1{j_1}^2 \end{split}\end{equation*} Obviously this expression is simplified if $I_1=I_2$, since then $I=I_2-I_1=0$. This choice of $I_1$ gives the first part of~\rf{41} the same form as the kinetic energy of a spherical ball moving in 3-space, so I call this choice the \defemph{ spherical gauge}. In the spherical gauge the flow lines of the reduced systems are trivial to determine: they are the intersection of the planes $(\kappa/2)w_2+a\pi_3=\mbox{constant}$ with the surfaces of revolution in Figure~\rf{200}. In the case of a nonsingular reduced space of nonzero dimension the flow is that is that of the flow on a 2-sphere with two stable equilibrium points. For the singular reduced spaces $-j_2=j_1>0$ the type of flow depends on whether or not the plane cuts through the cone at the $\pi_3=j_1$ singularity of~${w_2}^2=8(\pi_3+j_1)(\pi_3-j_1)^2$. That cone is \begin{equation*} |w_2|=4\sqrt{j_1}(j_1-\pi_3)\quad\Leftrightarrow\quad \pi_3= j_1-\frac1{4\sqrt{j_1}}|w_2|, \end{equation*} and so there is an unstable equilibrium at the singularity if the line $(\kappa/2)w_2+a\pi_3=aj_1$ passes through that cone, which is when $a^2<4\kappa^2j_1$, and a stable equilibrium otherwise. The former case is the flow on a 2-sphere with two stable equilibrium points and a single homoclinic connection to a singular point, while in the latter case this flow has two stable equilibrium points, one of which resides at a singularity. {\em For the remainder of this section I will assume the spherical gauge.} The equilibria of~\rf{54} correspond to relative equilibria for just the action of~$SO(3)$, since the space $\{(\pi,x)\}$ is the Poisson phase space of $P_{\mathrm{drift}}$ reduced by the $SO(3)$ action. These have a greater significance than relative equilibria which also use the $S^1$ normal form symmetry, since the full unapproximated Hamiltonian has the $SO(3)$ symmetry but it is only the truncated approximation which has the $SO(3)$ symmetry \emph{and} the normal form symmetry. As well, it is expedient to separate the equilibria of~\rf{54} at the outset of the analysis. Setting the right side of~\rf{54} to zero immediately gives $\pi_1=\pi_2=0$ and then $x_1\pi_3=x_2\pi_3=0$, so there are the following solutions: \begin{align} &x=0,\quad\pi=0\lb{57}\\ &x=0,\quad\pi_1=\pi_2=0,\quad\pi_3\ne 0\lb{58}\\ &x\ne 0,\quad\pi=0\lb{59} \end{align} The set of relative equilibria given by~\rf{59} contains relative equilibria equivalent under the $S^1$ normal form symmetry. The $SO(3)$ generator corresponding to~\rf{59} is, by substitution into~\rf{55}, $(\kappa x_1,\kappa x_2,0)$. The zero section of $T^*SO(3)$ (i.e. the set $\pi=0$) corresponds to perturbations of the original relative equilibrium $p_e$ such that the {\em perturbation} has zero total angular momentum, and in the nonresonant case such perturbations would imply no drift and no drift has been numerically observed for such. For the resonant case, however, one expects regular rotation around the axis $(\kappa x_1,\kappa x_2,0)$, which is perpendicular to the generator $\xi_e$. Nonzero $x$ corresponds to some ``reduced excitation''. Consequently, {\em at zero total angular momentum the system drifts so that the generator of the relative equilibrium moves along a fixed great circle at a rate dictated by a reduced excitation.} The relative equilibria~\rf{59} are the relative equilibria with nongeneric momenta {\em for the system}~\rf{41}. Consequently, the methods of~\ct{PatrickGW-1995.1} might be applied to them: the drifting relative equilibria could themselves drift. I will not, however, pursue this aspect here. To obtain a list of nonequivalent relative equilibria (i.e. a list of relative equilibria, no two of which are in the same $SO(3)\times S^1$ orbit), assume $A=\Id$ and discard Equation~\rf{55}, and, using the $S^1$ normal form symmetry, assume $\pi_2=0$ and $\pi_1\ge0$, and when $\pi_1=0$ assume that $x_2=0$. Equating the first of~\rf{54} with the $S^1$ infinitesimal generator of $s_e\in \mathbb R$ gives \begin{equation}\lb{62} \left[ \begin{array}{c} -\kappa x_2\pi_3+a\pi_2\\ \kappa x_1\pi_3-a\pi_1\\ \kappa(\pi_1x_2-\pi_2x_1)\\ \kappa\pi_2\\-\kappa\pi_1\end{array} \right] =\left[\begin{array}{c}-s_e\pi_2\\s_e\pi_1\\0\\-s_ex_2\\s_ex_1 \end{array}\right].\end{equation} Putting $\pi_2=0$ into the fourth component gives $s_ex_2=0$. Now $s_e=0$ corresponds to the $SO(3)$ relative equilibria~\rf{57}--\rf{59}, and in~\rf{57} and~\rf{58}, we have $x_2=0$, while in~\rf{59} we can assume $x_2=0$ and $x_1>0$ by the normal form symmetry. So if $s_e=0$ then we assume $x_2=0$, while if $s_e\ne0$ then $x_2=0$ anyway by $s_ex_2=0$. Putting $x_2=\pi_2=0$ into~\rf{62} then gives \begin{equation*} -\kappa\pi_1=s_ex_1,\qquad \kappa x_1\pi_3-a\pi_1=s_e\pi_1, \end{equation*} and these equations are easily solved to obtain the following list of nonequivalent relative equilibria: \begin{align} &\pi_1=\pi_2=x_1=x_2=0,\eta_e=(a+s_e)\mathbf k,\;\pi_3\in\mathbb R;\lb{64}\\ \begin{split}&\pi_2=x_2=0, \pi_3=\frac{\pi_1(ax_1-\kappa\pi_1)}{\kappa {x_1}^2}, s_e=-\frac{\kappa\pi_1}{x_1},\\ &\qquad\qquad\qquad\eta_e=\kappa x_1\mathbf i+(a+s_e)\mathbf k,\;\pi_1\in \mathbb R,x_1>0.\end{split}\lb{65} \end{align} Here the $SO(3)$ generators $\eta_e$ of these relative equilibria have been determined by comparison of~\rf{55} and the infinitesimal generator of the $SO(3)$ action of $P_{\mathrm{drift}}$ at the relative equilibrium. In~\rf{64} $\eta_e$ and $s_e$ are not unique due to the presence of isotropy in the phase space $P_{\mathrm{drift}}$. For later use, the characteristic polynomials of the linearizations of these relative equilibria are, respectively, \begin{equation*} p_1\equiv x^2(x^2+|\eta_e|^2) \bigl(x^4+(a^2+2as_e+2s_e^2-2\pi_3\kappa^2)x^2+ (s_e^2+as_e+\pi_3\kappa^2)^2\bigr) \end{equation*} and \begin{equation*} p_2\equiv \frac1{{x_1}^2}x^4(x^2+|\eta_e|^2)\bigl({x_1}^2x^2+\kappa^2{x_1}^4 +(ax_1-2\kappa\pi_1)^2\bigr). \end{equation*} For an understanding of the relative equilibria and the flow in general it is important to understand which relative equilibria reside on which reduced spaces. For~\rf{64}, $j_1=|\pi_3|$ and $j_2=\pi_3$, so for $\pi_3\le0$ one has $j_1=j_2$, which corresponds to pointlike reduced spaces, while for $\pi_3>0$ one has $j_1=-j_2$, which corresponds to the singular reduced spaces, and the relative equilibria occupy the singularities. As for~\rf{65}, substituting~$\rf{65}$ into $j_1$ and $j_2$ and eliminating $\pi_1$ gives \begin{multline}\lb{201} (3{x_1}^4+8j_2{x_1}^2+4{j_2}^2-4{j_1}^2)^2\kappa^2\\ +4a^2{x_1}^2({x_1}^2+2j_2-2j_1)({x_1}^2+2j_2+2j_1)=0 \end{multline} so it is a matter of solving this quartic in ${x_1}^2$ for its positive roots. The pointlike reduced spaces corresponding to $j_1=0$ are occupied by the relative equilibria of~\rf{65} after putting $\pi_1=0$, corresponding to the solutions $j_1=0$, $x_1=\sqrt{-2j_2}$ of~\rf{201}. \subsection{Partial reduction by the normal form symmetry} In this section I symplectically reduce the Hamiltonian system by the normal form symmetry at a value, say $\sigma$ of the momentum mapping $J^{\mbox{\scriptsize nf}}$. My objective is an interpretation of the reduced system via the equations of motion on the reduced space. Use the notation $r^2={x_1}^2+{x_1}^2$. The $\sigma$-level of the momentum map $J^{\mbox{\scriptsize nf}}$ is \begin{equation*} \pi_3=-\sigma-\frac12r^2. \end{equation*} Below, when $x$ appears in the context of a vector in $\mathbb R^3$, it is as $(x_1,x_2,0)$. The map \begin{equation}\lb{80} \Psi(A,\pi,x)\equiv(y,\alpha,z)\equiv \bigl(A\mathbf k,A(\pi-(\pi\cdot\mathbf k)\mathbf k),Ax\bigr) \end{equation} is a projection of the phase space $T^*SO(3)\times\mathbb R^2$ to the Whitney direct sum $T^*S^2\oplus TS^2$. The restriction of this map to the $\sigma$ level set of $J^{\mbox{\scriptsize nf}}$ is clearly a quotient map for the normal form action, and thus the symplectic reduced space is $T^*S^2\oplus TS^2$. The action of $SO(3)$ on $T^*S^2\oplus TS^2$ becomes \begin{equation*} A\cdot(y,\alpha,z)=(Ay,A\alpha,Az) \end{equation*} To find the reduced vector field at $(y,\alpha,z)\in T^*S^2\oplus TS^2$, first evaluate the original Hamiltonian vector field, namely \begin{alignat}{2} & \Omega\equiv I_1\pi+I((\pi\cdot\mathbf k)+a)\mathbf k,&\qquad &A^{-1}\frac{dg}{dt}=\Omega^\wedge,\nonumber\\ &\frac{d\pi}{dt}=\dot\pi=\pi\times\Omega+a\mathbf k, &\qquad\qquad&\frac{dx}{dt}=\dot x=-\kappa\mathbf k\times\pi\lb{81} \end{alignat} at the point \begin{equation} \pi= A^{-1}\alpha-\left(\sigma+\frac12r^2\right)\mathbf k,\quad x=A^{-1}z,\lb{82} \end{equation} where $A$ is chosen so that $A\mathbf k=y$, and then apply the derivative of~\rf{80}, which is \begin{equation}\begin{split} \lefteqn{T\Psi(A,\pi,x,\Omega,\dot\pi,\dot x)}\\ &\equiv \dbyd t0\Psi\bigl(A\exp(\Omega^\wedge t),\pi+t\dot\pi,x+t\dot x\bigr)\\ &=A\cdot\bigl(\Omega\times\mathbf k, \Omega\times(\pi-(\pi\cdot\mathbf k)\mathbf k)+\dot\pi -(\dot\pi\cdot\mathbf k)\mathbf k, \dot x+\Omega\times x\bigr). \end{split}\lb{83}\end{equation} Thus it is a matter of substituting~\rf{81} and~\rf{82} into~\rf{83}. Without care the calculation can be onerous; however, it is straightforward, and yields the equations of motion \begin{gather} \frac{dy}{dt}=(I_1\alpha+\kappa z)\times y,\lb{841}\\ \frac{d\alpha}{dt}=-\pi_3(I_1\alpha+\kappa z)\times y +\kappa z\times \alpha,\lb{842}\\ \frac{dz}{dt}=\big(\kappa\alpha-(I_2\pi_3+a)z\bigr)\times y +I_1\alpha\times z,\lb{843} \end{gather} where $\pi_3$ stands for $-(\sigma+|z|^2/2)$. To these equations must be added the constraints $|y|=1$, $\alpha\cdot y=0$ and $z\cdot y=0$. I want to impose the viewpoint that there is a ``particle'' at~$y$ having direction~$z$. To do this, I use~\rf{841} to replace $\alpha$ with $dy/dt$ in~\rf{842} and~\rf{843} while writing equation~\rf{842} as a second order equation in~$y$. I also use the standard Levi-Cevita connection of $S^2$ for the time derivatives. Thus, Equation~\rf{842} becomes \begin{equation*}\begin{split} \frac{d^2y}{dt^2}&=(I_1\alpha+\kappa z)\times\frac{dy}{dt}+ \left(I_1\frac{d\alpha}{dt}+\kappa\frac{dz}{dt}\right)\times y\\ &=\bigl(I_1\alpha+\kappa z\bigr)\times\bigl((I_1\alpha+\kappa z) \times y\bigr)\\ &\qquad\mbox{}+I_1\pi_3(I_1\alpha+\kappa z)-\kappa\bigl(\kappa\alpha -(I_2\pi_3+a)z\bigr)\\ &=-\left|\frac{dy}{dt}\right|^2y+I_1\pi_3y\times\frac{dy}{dt} -\kappa\left(\frac\kappa{I_1} \left(y\times\frac{dy}{dt}-\kappa z\right)-(I_2\pi_3+a)z\right), \end{split}\end{equation*} and similarly with Equation~\rf{843}, so that \begin{gather} \frac{\nabla^2y}{dt^2}=\frac{{I_1}^2\pi_3-\kappa^2}{I_1}y\times\frac{dy}{dt} +\kappa\left(\frac{\kappa^2}{I_1}+I_2\pi_3+a\right)z\lb{85}\\ \frac{\nabla z}{dt}=\frac\kappa{I_1}\frac{dy}{dt} +\left(\frac{\kappa^2}{I_1}+I_2\pi_3+a\right)y\times z.\lb{86} \end{gather} Some aspects of the particle become apparent by replacing $\alpha$ with $dy/dt$ in the total energy and angular momentum. Using~\rf{82}, the total energy is \begin{equation}\begin{split} H&=\frac12|\alpha|^2+\frac12{\pi_3}^2+a\pi_3+\kappa z\cdot\alpha\\ &=\frac1{2I_1}|I_1\alpha+\kappa z|+\frac12{\pi_3}^2+a\pi_3 -\frac{\kappa^2}{2I_1}|z|^2\\ &=\frac{1}{2I_1}\left|\frac{dy}{dt}\right|^2 +\frac{1}{2I_2}\left(\frac{\kappa^2}{I_1}+I_2\pi_3+a\right)^2 +\frac{\kappa^2\sigma}{I_1},\end{split}\lb{93} \end{equation} so the internal energy of the particle is \begin{equation*} E_{\mbox{\scriptsize int}} =\frac{1}{2I_2}\left(\frac{\kappa^2}{I_1}+I_2\pi_3+a\right)^2. \end{equation*} The total angular momentum, or the $SO(3)$-momentum-mapping is, using~\rf{82}, \begin{equation} J=A\pi=\alpha-\pi_3\mathbf k=\frac1{I_1}y\times\frac{dy}{dt} -\frac{\kappa}{I_1}z-\pi_3y,\lb{92} \end{equation} and below I replace $z$ with the tangent to the sphere part of the angular momentum, which by~\rf{92} is \begin{equation*} L\equiv-\frac\kappa{I_1}z, \end{equation*} while the total angular momentum attributed to the particle must be $L+\pi_3y$. By the first terms of both~\rf{93} and~\rf{92}, the particle should be viewed as having mass $1/I_1$. From~\ct{JacksonJD-1975.1} the classical nonrelativistic equations of motion on $|y|=1$ of a particle of mass $m$, charge $Q$, and gyromagnetic ratio $\Gamma$, and magnetic moment $\mathbf m=\Gamma L$, are \begin{gather} m\frac{\nabla^2y}{dt^2}= \frac QcB\times \mathbf v+\nabla(\mathbf m\cdot B)-(\nabla\cdot B)\mathbf m,\lb{87}\\ \frac{\nabla L}{dt}={\mathcal I}+\Gamma L\times B,\lb{88} \end{gather} where $c$ is the velocity of light and $\mathbf v$ is the particle's velocity. Here $\mathcal I$ is compensates for the inductive forces that would, were $\mathbf m$ to be constant, be required to maintain the constant current that generates~$\mathbf m$ itself. One can fit equations~\rf{85} and~\rf{86} to these equations as follows. If one takes $B=y$, comparison of the first terms of~\rf{86} and~\rf{88} gives the inductive term \begin{equation} {\mathcal I}=-\frac{\kappa^2}{{I_1}^2}\frac{dy}{dt},\lb{95} \end{equation} while the second term of~\rf{86} matches the second term of~\rf{88} if the gyromagnetic ratio is \begin{equation} \Gamma=\frac{{I_1}^2}{\kappa^2}\left(\frac{\kappa^2}{I_1}+I_2\pi_3 +a\right).\lb{89} \end{equation} Matching the second term of~\rf{85} with the third term of~\rf{87} implies $\Gamma$ is $\kappa^2/3I_1$ times the left side of~\rf{89}. So choose a gauge so that $I_1=\kappa^2/3$. Then matching the first term of~\rf{85} with the first term of~\rf{87} yields \begin{equation*} \frac Qc=\pi_3-\frac{\kappa^2}{{I_1}^2}. \end{equation*} Finally, I note that the inductive term~\rf{95} is exactly what is required to balance the rate of change of internal energy with the work done on the particle through its magnetic moment via the last term of~\rf{85}. \section{Example: Two coupled rods} In this section I illustrate and verify the above theory using numerical simulations of the system of two axially symmetric rods which are joined by a frictionless ball-and-socket joint. In the center of mass frame this system can be cast as a geodesic flow with configuration space $SO(3)^2=\{(A_1,A_2)\}$ and kinetic energy metric \begin{equation*} L=\frac{1}{2}\left[\begin{array}{cc}{\Omega_1}^t& {\Omega_2}^t\end{array}\right]J(A)\left[\begin{array}{c}\Omega_1\\ \Omega_2 \end{array}\right], \end{equation*} where $A=[A^{ij}]={A_1}^tA_2$, $\Omega_1$ and $\Omega_2$ are body referenced (left translation) angular velocities, and \begin{equation*}J(A)=\left[\begin{array}{cccccc} 1&0&0&-\beta A^{22}&\phantom{-}\beta A^{21}&0\\ 0&1&0&\phantom{-}\beta A^{12}&-\beta A^{11}&0\\ 0&0&\alpha&0&0&0\\ -\beta A^{22}&\phantom{-}\beta A^{12}&0&1&0&0\\ \phantom{-}\beta A^{21}&-\beta A^{11}&0&0&1&0\\ 0&0&0&0&0&\alpha\end{array}\right]. \end{equation*} Here $0<\alpha<2-2\beta$ is a parameter which increases with the diameter of the rods and $0\le\beta<1$ is a parameter which measures the degree of coupling between the rods; $\beta$ is zero if the joint lies at the mutual centers of mass of the rods. The continuous symmetry group of this system is $SO(3)\times (S^1)^2$, which acts on the configuration space $SO(3)^2$ by \begin{equation*} (B,\theta _1,\theta _2 )\cdot(A_1,A_2)=\bigl(BA_1\exp(-\theta_1 \mathbf k^\wedge),BA_2\exp(-\theta_2\mathbf k^\wedge)\bigr). \end{equation*} All relative equilibria for this system are explicitly known, as are their formal stability; a complete list can be found in~\ct{PatrickGW-1991.1}. The relative equilibria of interest here are the phase space points $\bigl(\Id,\exp(\theta\mathbf j^\wedge),\Omega_1,\Omega_2\bigr)$ parameterized by~$t_1,t_2,\theta\in\mathbb R$ such that $t_1\ne0,t_2\ne0, 0<\theta<\pi$, where \begin{equation*} \Omega_1=t_1\mathbf i-\kappa^{t_1t_2\theta}_1\mathbf k, \Omega_2=t_2\mathbf i+\kappa^{t_2t_1\theta}_1\mathbf k \end{equation*} and where \begin{equation*} \kappa^{t_1t_2\theta}_\gamma\equiv\frac{(t_1\cos \theta-t_2)(\beta t_2-\gamma t_1)} {\alpha t_1\sin \theta}. \end{equation*} The corresponding generators $(\Omega,\sigma_1,\sigma_2)\in so(3)\times\mathbb R^2$ are given by \begin{equation*} \sigma_1=\kappa^{t_1t_2\theta}_{1-\alpha},\sigma_2=- \kappa^{t_2t_1\theta}_{1-\alpha},\Omega=t_1\mathbf i+\frac{t_1\cos \theta-t_2} {\sin \theta}\mathbf k. \end{equation*} By calculating the linearizations, one sees that a 1:1 group reduced resonance with zero total-angular-momentum can be arranged by setting \begin{equation*} \beta=\frac{2t_1t_2}{{t_1}^2+{t_2}^2}, \quad 3{t_1}^4+{t_1}^2{t_2}^2(4\cos^2\theta-10)+3{t_2}^4=0, \end{equation*} and for the work presented in this section I have chosen to perturb the particular relative equilibrium obtained by setting \begin{equation}\lb{900} \theta=\frac\pi3,t_1=1,t_2=\frac{1+\sqrt 5}{2}, \alpha=\frac12,\beta=\frac2{\sqrt 5}. \end{equation} \subsection{Calculations for the coupled rod system} To compare the dynamics of the drift system to the dynamics of the coupled rod system near the relative equilibrium~\rf{900}, it is necessary to calculate the splitting~\rf{1} or~\rf{2} in this special case. Here is one general way, inspired by the proof of Theorem~3.1.19 of~\ct{AbrahamRMarsdenJE-1978.1}: \begin{enumerate}\renewcommand{\labelenumi}{(\alph{enumi})} \item Calculate the semisimple part $S_{p_e}$ and the nilpotent part $N_{p_e}$ of the linearization $dX_{H_{\xi_e}}(p_e)$. \item Since $S_{p_e}$ is semisimple it has a basis of eigenvectors, some complex and some real, and the complex eigenvectors may be grouped into complex conjugate pairs. Taking the real and imaginary parts of one eigenvector in each pair, and then including the real eigenvectors yields a basis $\mathcal B$ of $T_{p_e}P$ in which $S_{p_e}$ is {\em skew symmetric}. \item Let the matrix of $\omega(p_e)$ with respect to the basis $\mathcal B$ be $-W$. Since $S$ is infinitesimally symplectic and skew, $0=S^tW+WS=-SW+WS$, so $S$ and $W$ commute. By going to a basis of eigenvectors of the positive matrix $-W^2$, find a symmetric, positive square-root $B$ of $-W^2$ such that $B$ and $S$ commute, and set $J=WB^{-1}$. Using the basis $\mathcal B$, regard $J$ as its corresponding operator on $T_{p_e}P$, and regard $B$ as the bilinear form on $T_{p_e} P$ corresponding to $(x,y)\mapsto x^tBy$. Then \begin{equation*} \omega(p_e)(v,w)=B(Jv,w),\quad J^2=-\Id, \end{equation*} and $J$ commutes with $S_{p_e}$. \item Set $Z=J(\mathfrak g_{\mu_e})$, so $Z$ is a $S_{p_e}$ invariant Lagrangian complement to $\mathfrak g_{\mu_e}$ in $\mathfrak g_{\mu_e}\oplus Z$. Choose an $\onm{Ad}$-invariant complement $\mathfrak b$ to $\mathfrak g_{\mu_e}$, and set $W_{\!\mathrm{red}}=(\mathfrak g_{\mu_e} p_e\oplus Z\oplus\mathfrak bp_e)^{\omega\perp}$. Then $W_{\!\mathrm{red}}$ is symplectic, $S_{p_e}$ invariant and is contained in $\onm{ker} dJ(p_e)$, since $W_{\!\mathrm{red}}\subseteq \mathfrak gp_e^{\omega\perp}$. The subspace $W_{\!\mathrm{red}}$ complements $\mathfrak g_{\mu_e}$ in $\onm{ker}dJ(p_e)$ since $\mathfrak g_{\mu_e}\subseteq W_{\!\mathrm{red}}^{\omega\perp}$ and $W_{\!\mathrm{red}}$ is symplectic. \end{enumerate} In the case of two coupled rods, further refinement of the splitting of $W_{\!\mathrm{red}}$ to resonant and nonresonant parts is not necessary, since the reduced spaces for zero total-angular-momentum are two dimensional; there is no ``nonresonant'' part of the phase space. Final adjustments to achieve the spherical gauge are easily arranged using~\rf{12} and~\rf{13}. The end result of all this is a basis of the tangent space at the relative equilibrium which reflects the splitting~\rf{2}, a basis of the Lie algebra $so(3)\times\mathbb R^3$, reflecting the splitting~\rf{102} into resonant and nonresonant parts, the information \begin{equation}\lb{3000} \kappa=.9115064,\quad I_1=I_2=4.321619, \end{equation} and the information that~\rf{900} is a ``$+$''~type relative equilibrium. Further to calculating the splitting~\rf{1} or~\rf{2}, means must be found to translate initial conditions of the drift system to perturbations (i.e. initial conditions near~\rf{900}) of the coupled rod system. For this a map was used, say $\Psi$, of the tangent space $T_{p_e}P$ into phase space $P$ such that $T_{p_e}\Psi$ is the identity at $p_e$. The map $\Psi$ only injected initial conditions of the drift system into perturbations of~\rf{900} to first order; actual correspondence could not be easily achieved due to the implicit nature of the coordinates provided by the isotropic embedding theorem. Moreover, $\Psi$ did not match the momentum of the drift system exactly to a momentum perturbation of the actual system, while practice indicated that an exact match of these momenta is important. For the comparisons just below, given an initial condition of the drift system with a particular momentum, an initial condition of the coupled rod system with matching momentum was obtained as $\Psi$ of another nearby initial condition of the drift system. That nearby initial condition of the drift system was obtained by a slight (second order) iterative refinement of the original initial condition of the drift system. By the conventions of Section~4, and since the relative equilibrium~\rf{900} is of ``$+$''~type, the vector $(\pi_1,\pi_2,\pi_3)=\mathbf k$ is parallel to the normalized rotation vector $\xi_e$ of the coupled rod system. Also, the quantity $A\in SO(3)$ of the drift system corresponds to the drifted orientation of the coupled rod system. Thus, the quantity $A\mathbf k$ of the drift system corresponds to the drifting rotation vector of the coupled rod system. Now, the rotation vector of the relative equilibrium~\rf{900} is a constant linear sum of the locations of the rods, exactly because the rods when in the relative equilibrium rotate around that vector: if one defines \begin{equation*} \tau\equiv \frac{1}{\sin\theta}\left(t_1A_1\mathbf k -t_2A_2\mathbf k\right) \end{equation*} then $\tau=\xi_e$ at the relative equilibrium~\rf{900} and also at any reorientation of~\rf{900}. Thus, for a perturbation of~\rf{900}, $\tau$ approximates the drifting rotation vector and the prediction of the preceding theory is that \begin{equation*} \bar\tau\equiv\frac\tau{|\tau|}\approx A\mathbf k. \end{equation*} The numerical verifications I have undertaken consist of predictions of the drift system for the motion of $A\mathbf k$ and the comparison of these predictions with evolution of the the unit vector $\bar\tau$ calculated via a symplectic integration of the coupled rod system. The particular algorithm used was a (implicit) Riemannian leapfrog algorithm~(\ct{PatrickGW-1996.1}). \begin{figure}\setlength{\unitlength}{1in}\centering \begin{picture}(4.5,2) \put(1.,1){\makebox(0,0){\epsfbox{fig4a.eps}}} \put(3.4,1){\makebox(0,0){\epsfbox{fig4b.eps}}} \end{picture} \caption{\lb{2500}\protect\footnotesize\it Perturbations within zero-total-angular momentum. Left: the paths traced out by $\bar\tau$, resulting from $6$~initial conditions within zero total-angular-momentum. Right: the rotation rates of $\tau$ vs. the magnitude of the perturbation $|x|$.} \end{figure} \subsection{Numerical results} \subsubsection{First comparison: zero total-angular-momentum} By the results in Section~(4.1), the part of the drift phase space corresponding to perturbations with zero total-angular-momentum is occupied by $SO(3)$ relative equilibria, and the motion of $A\mathbf k$ from initial condition $\pi=0$ and $x$ arbitrary is that of uniform rotation at angular frequency $\kappa|x|$ along a great circle through $\mathbf k$ and perpendicular to $(x_1,x_2,0)$. Translated to the coupled rod system, the prediction is that $\bar\tau$ undergoes regular rotation in a great circle through $\bar\tau(0)$, the particular great circle regularly rotating as $x$ is rotated. The left of Figure~\rf{2500} shows the motion of $\tau$ as $x$ is varied from $0$ to $0$ through $5\pi/6$ radians by increments of $\pi/6$, so the coupled rod system conforms to this prediction. As the magnitude of $x$ is varied the great circles should be traced out at an angular rotation rate $\kappa |x|$, so $\kappa$ can be determined by plotting that rotation rate against $|x|$; this is done on the right of Figure~\rf{2500}. The rotation rates fit well to the curve \begin{equation*} \mbox{Rate}= .9284|x|+.7750|x|^2, \end{equation*} so from the simulation $\kappa$ is $.9284$, which is $1.8\%$ off the calculated value of $.9115064$ already displayed in~\rf{3000}. \subsubsection{Second comparison: a stable relative equilibrium of the drift system} For another comparison, I examined the coupled rod system for perturbations of~\rf{900} corresponding to being near one of the stable relative equilibria in the list~\rf{65}. For perturbations of~\rf{900} corresponding to the values \begin{equation}\lb{1001} \pi_1=.001,\quad x_1=.04 \end{equation} in~\rf{65} the predictions of the drift system are that $\tau$ moves as small periodic oscillation of the full reduction of the drift system superimposed on the periodic motion $\exp(\eta_e^\wedge t)\mathbf k$, where $\eta_e$ is the $SO(3)$ generator in the list~\rf{65}. The two predicted frequencies are the linearized frequency of the corresponding equilibrium on the reduced space of the drift system and the rotation frequency of $\eta_e$, which are by substitution of~\rf{1001} into~\rf{65} and/or~\rf{201}, respectively, \begin{equation}\lb{2004} .04477,\quad .05837. \end{equation} Again by substitution of~\rf{1001} into~\rf{65}, the motion of $\tau$ should be such that its projection onto the unit vector $\eta_e/|\eta_e|$ is near $-.5300$. \begin{figure}[p]\setlength{\unitlength}{1in}\centering \begin{picture}(4.5,2) \put(1.,1){\makebox(0,0){\epsfbox{fig32a.eps}}} \put(3.4,1){\makebox(0,0){\epsfbox{fig32b.eps}}} \end{picture} \vspace*{.05in}\begin{picture}(4.5,2) \put(2.25,1){\makebox(0,0){\epsfbox{fig34.eps}}} \end{picture} \vspace*{.15in}\footnotesize\begin{tabular}{rrrrrr|rrrrrr} Power&Freq.&Freq.&$z_1$&$z_1$&$z_3$& Power&Freq.&Freq.&$z_1$&$z_1$&$z_3$\\ \hline 1.195 & .04449 & .04449 & 1 & 0 & 0 & 8.766 & .2266 & .2266 & 0 & 1 & 1 \\ 2.935 & .05791 & .05791 & 0 & 1 & 0 & 8.848 & .1822 & .1822 & -1 & 1 & 1 \\ 3.485 & .01342 & .01342 & -1 & 1 & 0 & 8.874 & .1549 & .1553 & 1 & -1 & 1 \\ 4.871 & .1024 & .1024 & 1 & 1 & 0 & 8.893 & .1607 & .1603 & 1 & 2 & 0 \\ 5.812 & .1687 & .1687 & 0 & 0 & 1 & 10.3 & .1296 & .1292 & -1 & 3 & 0 \\ 6.146 & .1239 & .1243 & -1 & 0 & 1 & 10.58 & .1741 & .1737 & 0 & 3 & 0 \\ 6.485 & .07133 & .07133 & -1 & 2 & 0 & 10.63 & .2711 & .2711 & 1 & 1 & 1 \\ 6.817 & .1162 & .1158 & 0 & 2 & 0 & 12.25 & .2401 & .2401 & -1 & 2 & 1 \\ 7.102 & .2132 & .2132 & 1 & 0 & 1 & 12.36 & .2846 & .2846 & 0 & 2 & 1 \\ 8.24 & .06596 & .06634 & -1 & -1 & 1 & 12.91 & .3371 & .3375 & 0 & 0 & 2 \\ 8.742 & .1104 & .1108 & 0 & -1 & 1 & 12.98 & .2926 & .293 & -1 & 0 & 2 \\ \end{tabular} \vspace*{.15in}\footnotesize \caption{\lb{1000}\protect\footnotesize\it Some results of the simulation of the coupled rod system for a perturbation corresponding to a elliptic relative equilibrium of the drift system. Top left: the evolution of the rotation vector projected onto the unit sphere. Top right: the time evolution of the height corresponding to the top left. Middle: the power spectrum of a signal derived from the drifting motion; note the vertical logarithmic scale. Bottom table: the peaks in the power spectrum, sorted by power. The second column tabulates the frequencies of the peaks and the third column tabulates the harmonics $z_1A+Z_2B+Z_3C$ of the frequencies $A$, $B$ and $C$ indicated on the frequency axis of the power spectrum.} \end{figure} Figure~\rf{1000} shows the results of a simulation of the coupled rod system for the perturbation of~\rf{900} corresponding to the values~\rf{1001}. In the simulation the normalized rotation vector $\bar\tau$ moved on the unit sphere as shown in the top left of the Figure; the sense of the rotation is clockwise as seen from the top, so that $\eta_e$ is pointing down and away from you as you look upon the Figure. The height of the rotation vector has been graphed in the top right of Figure~\rf{1000}, and is visibly quasiperiodic with about two frequencies. The measured average height of the motion was $.4970$, compared with the prediction of $.5300$. In the middle of Figure~\rf{1000} is the power spectrum of one of the horizontal components of $\tau$. As is shown in the table immediately below the power spectrum, the power spectrum is consistent with the 3 fundamental angular frequencies \begin{equation}\lb{2005} .04449,\quad .05791,\quad .1687, \end{equation} all other peaks being harmonics of these. The first two of these are by far the largest peaks (note the vertical logarithmic scale in the power spectrum) and they agree with the the predicted frequencies~\rf{2004}. The third observed frequency in~\rf{2005} corresponds to a far less prominent peak and is likely a frequency associated to the higher order terms that have been truncated in the drift approximation. To give an idea of the speed of the drift, on the short simulation corresponding to the top left and right of Figure~\rf{1000}, the system rotated about 300 times while $\bar\tau$ rotated about 7~times. Thus the drift was about 50 times slower than the original relative equilibrium. Due to this, these kinds of simulations can be long time: over the duration of the simulation that generated the power spectrum in Figure~\rf{2} the rods rotated about 10,000 times. \subsubsection{Third comparison: a singularity-induced phase jump.} As shown in Section~(4.2), the singularities of the $-j_1=j_2$ reduced phase spaces of the drift system are occupied by an unstable equilibria corresponding to $SO(3)$~relative equilibria of the drift system. A single homoclinic orbit, tracing out the intersection of each reduced space with the plane $w_2=0$, emanates from every such equilibrium. These homoclinic orbits are interesting features of the drift system, in part because they provide an opportunity to investigate dynamics near singularities of reduced spaces. I will show here that, due to the singularity of the reduced space, the drift system suffers a jump in its reconstruction phase as initial conditions traverse the homoclinic orbits, and (numerically) that this feature of the drift system persists to the coupled rod system. The reconstruction phase jump shows the presence of Hamiltonian monodromy in the completely integrable drift system; for more information on Hamiltonian monodromy see~\ct{CushmanRHBatesLM-1997.1}, page~175, the summary on page~403 of the same reference, as well as~\ct{BatesLM-1991.1}. So choose $j_1$ small and consider perturbations of the relative equilibrium \begin{equation*} \pi=j_1\mathbf k,\quad x=0,\quad A=\Id, \end{equation*} where by abuse of notation $j_1$ serves as the constant value of the Casimir~\rf{70} of the same name. This relative equilibrium corresponds to the equilibrium \begin{equation} \lb{550}w_1=w_2=0,\quad \pi_3=j_1 \end{equation} on the $SO(3)\times S^1$~reduced space. Near~\rf{550} the motion on the reduced space of the drift system is periodic along the intersection of the $w_2=h$ plane where $h$ is small: for a long time the system remains near the singularity, and then moves off, passes near $\pi_3=j_1$ and then returns to the singularity, similar to, for example, the motion of an inverted pendulum. One $S^1$ reconstruction gives the motion on the phase space $\{(x,\pi)\}$ and, since the motion is periodic on the $SO(3)\times S^1$~reduced space, there is a well defined \defemph{$S^1$ reconstruction phase}. On the phase space $\{(\pi,x)\}$, the only points on the level sets of $j_1$ and $j_2$ that map to~\rf{550} are \begin{equation}\lb{551} \pi=j_1\mathbf k,\quad x=0, \end{equation} since $\pi_3=j_1=-j_2$ implies by~\rf{70} and~\rf{71} that $\pi_1=\pi_2=x_1=x_2=0$. Thus, when the $SO(3)\times S^1$ reduced system is near~\rf{550}, the reduced system with phase space~$\{(x,\pi)\}$ is near~\rf{551}. An $SO(3)$ reconstruction gives the motion on the phase space $\mathbb R\times T^*SO(3)=\{(x,A,\pi)\}$, starting, say, at $A=\Id$. By conservation of the momentum $J=A\pi$, the variable $A$ is nearly a rotation about $\mathbf k$ whenever the reduced system is near~\rf{551}. Thus, the motion of the point $A\mathbf k$ on the unit sphere is this: \emph{for a long time $A\mathbf k$ remains near $\mathbf k$, then moves off, passes near $-\mathbf k$, and then returns near to $\mathbf k$. This motion repeats but rotated (with respect to the previous excursion from $\mathbf k$) by a excursion-independent angle about $\mathbf k$. This rotation is what I mean by the \defemph{$SO(3)$~reconstruction phase shift} of the motion.} \begin{figure}\setlength{\unitlength}{1in}\centering \begin{picture}(4.5,2) \put(1.,1){\makebox(0,0){\epsfbox{fig6.eps}}} \put(3.4,1){\makebox(0,0){\epsfbox{fig2a.eps}}} \end{picture} \caption{\lb{4002}\protect\footnotesize\it Left: the motion of $\bar\tau$ for a perturbation corresponding to being near the homoclinic orbit of the reduced space of the drift system. The vector $\bar\tau$ begins near the top of the sphere and moves off to the left, comes around the sphere, makes an approximate $\pi/2$ turn, then moves away from you, and goes around the sphere again. On the right: the $SO(3)$ reconstruction phase of the coupled rod system as initial conditions are rotated around the singularity of the reduced phase space of the drift system.} \end{figure} By simulation of the coupled rod system I have verified that the gross details of the motion of $A\mathbf k$ in the drift system also occur in the coupled rod system (see the left of Figure~\rf{4002}). This situation is robust; no particular care is required, in choosing the coupled rod system's initial conditions to evoke these kinds of motions. The robust nature under addition of higher order terms of the homoclinic orbit is expected through averaging theorems such as the one in~\ct{GuckenheimerJHolmesP-1983.1},~page~168. It is the behavior of the reconstruction phase of the drift system near to $\pi=0$ that is relevant for predictions of the motion of the coupled rod system near to its resonant relative equilibrium. One way to examine this behavior is to numerically integrate initial conditions of the drift system that are reconstructions of initial conditions starting at a small circle surrounding the singularity of the $SO(3)\times S^1$ reduced space. For the $S^1$ reconstruction phase shift the result is a constant phase shift of $\pm\pi/2$ with a jump of $\pi$ as the homoclinic orbit is traversed. The corresponding phase shift of $A\mathbf k$ in fact becomes undefined as the perturbation vanishes. The reason for this is that the equation of motion for $A$, namely Equation~\rf{55}, becomes, near to~\rf{551}, the equation \begin{equation*} \frac{dA}{dt}=(I_2j_1+a)\mathbf k^\wedge, \end{equation*} while $A(t)\mathbf k$ is not exactly $\mathbf k$ due the the presence of the perturbation itself. Consequently the $A\mathbf k$ picks up a rotation of angular frequency $I_2j_1+a$ during its long visit of~\rf{551}. As the perturbation vanishes this long visit becomes an eternity and this undefines the $SO(3)$ reconstruction phase shift as an asymptotic effect. However, viewed from a frame that corotates with the same angular frequency, namely $I_2j_1+a$, one can expect a well defined phase shift. I have numerically verified this: from the corotating frame the $SO(3)$ reconstruction phase shift of the drift system is nearly $\pm\pi/2$ with a jump of $\pi$ as the homoclinic orbit is traversed. This compares favorably with the $SO(3)$ reconstruction phase shift of the coupled rod system shown in right of Figure~\rf{4002}. Finally, so that the comparison of these phases of the drift system and the coupled rod system are not entirely numerical, I give a calculation showing that, asymptotically as the homoclinic orbit is approached, these phases arise mostly from the singularity of the reduced space. The $S^1$ reconstruction phase of the drift system may be calculated by a slight modification (to allow reparameterization) of the usual reconstruction method found in~\ct{AbrahamRMarsdenJE-1978.1}: Generally, suppose $H$ is a Hamiltonian on a symplectic phase space $P$ and that a curve $c_\mu(s)$ is a reparameterization an evolution $c_\mu(t)$ on some Marsden-Weinstien reduced phase space~$P_\mu=J^{-1}(\mu)/G_\mu$. Choose a curve $d(s)\in J^{-1}(\mu)$ such that $c=\pi_\mu\circ d$, where $\pi_\mu:J^{-1}(\mu)\rightarrow P_\mu$ is the quotient projection. Then there is a unique smooth curve $\xi(s)\in\mathfrak g_\mu$ and a unique smooth function $a(s)$ such that \begin{equation}\lb{554} X_H\bigl(d(s)\bigr)=\xi(s)p(s)+a(s)d^\prime(s). \end{equation} By differentiation, the curve $g\bigl(s(t)\bigr)d\bigl(s(t)\bigr)$ satisfies Hamilton's equations on $P$ if \begin{equation*} g^{-1}\frac{dg}{ds}=\frac1{a(s)}\xi(s),\quad\frac{dt}{ds}=\frac1{a(s)}. \end{equation*} Perfectly obvious generalizations hold if $d$ is just a $1$-manifold in $J^{-1}(\mu)$ covering the image of $c_\mu$. Particularly, by setting $\pi_2=0$, $d$ can be the subset of $P_{\mathrm{drift}}$ defined by \begin{gather} \pi_1^2+\pi_2^2=j_1^2,\quad \pi_2=0,\quad x_1=\frac{h}{2\pi_1},\quad x_2=\frac{-w_1}{2\pi_1},\lb{5553}\\ w_1^2+h^2=8(\pi_3^2-j_1^2)(\pi_3-j_1),\lb{553} \end{gather} in which case one calculates using~\rf{54},~\rf{554},~\rf{5553} and~\rf{553} that \begin{equation}\lb{555} \frac{d\theta}{d\pi_3}=\frac{-h\pi_3}{({j_1}^2-{\pi_3}^2)w_1},\quad \frac{d\pi_3}{dt}=-\frac{\kappa w_1}2 \end{equation} where $\theta$ is the $S^1$ reconstruction phase as the counterclockwise angle of the vector $\bigl(\pi_1(t),\pi_2(t)\bigr)$ from the $\pi_2=0$ axis. Let $r_1(h)<r_2(h)<r_3(h)$ be the roots of the cubic $8(\pi_3-j_1)^2(\pi_3+j_1)-h^2$. By standard perturbation arguments \begin{gather*} r_1=-j_1+\frac1{32{j_1}^2}h^2+O(h^3),\\ r_2=j_1-\frac1{4\sqrt{j_1}}h+O(h^2),\\ r_3=j_1+\frac1{4\sqrt{j_1}}h+O(h^2). \end{gather*} The $S^1$ reconstruction phase over the curve $w_1^2+h^2=8(\pi_3^2-j_1^2)(\pi_3-j_1)$ is calculated as follows. Let the reduced system start at $w_1=0$, $\pi_3=r_2$ at time $t=t_0$, move to $w_1=0$, $\pi_3=r_1$ at time $t=t_1$ and then complete its periodic orbit by moving back to $w_1=0$, $\pi_3=r_2$ at time $t_2$. Then the contribution of~\rf{555} to phase shift over the interval $[t_0,t_1]$ is \begin{equation*} \phi(t_1)-\phi(t_0)=\int_{t_0}^{t_1}\frac{d\phi}{dt}\,dt =\int_{\pi_3(t_0)}^{\pi_3(t_1)}\frac{d\phi}{dt}\frac{dt}{d\pi_3}\,d\pi_3 =\int_{r_2}^{r_1}\frac{d\phi}{d\pi_3}\,d\pi_3, \end{equation*} where the positive square root must be used when solving for $w_1$ in~\rf{553}, since $\pi_3$ must immediately decrease after time $t=t_0$, and by the second of~\rf{555}, $w_1$ is positive over the interval $[t_0,t_1]$. The contribution of~\rf{555} over the interval $[t_1,t_2]$ is identical (use the negative root here). Thus, the total phase shift $\phi\equiv\phi(t_2)-\phi(t_0)$ over the loop as the loop approaches the homoclinic orbit through $h>0$ is \begin{equation}\lb{557} \phi=-2\lim_{h\rightarrow0^+}\int_{r_1}^{r_2}\frac{-h\pi_3}{({j_1}^2-{\pi_3}^2) \sqrt{8(\pi_3-r_1)(r_2-\pi_3)(r_3-\pi_3)}}\,d\pi_3. \end{equation} For small $h$ the integrand of~\rf{557} is small away from its two singularities at $\pi_3=r_1$ and $\pi_3=r_2$. I begin with the left singularity at $\pi_3=r_1$, so I calculate \begin{equation*} \phi_1\equiv\lim_{h\rightarrow0^+}\int_{r_1}^0 \frac{-h\pi_3}{({j_1}^2-{\pi_3}^2) \sqrt{(\pi_3-r_1)(r_2-\pi_3)(r_3-\pi_3)}}\,d\pi_3. \end{equation*} Elementary estimates show that zero error is made as $h\rightarrow0$ by the replacement of \begin{equation*} \frac{\pi_3}{(j_1-\pi_3)\sqrt{(\pi_3-r_2)(\pi_3-r_3)}} \end{equation*} with its limit as $h\rightarrow0$ of its evaluation at $\pi_3=r_1$, which is \begin{equation*} \lim_{h\rightarrow0^+}\frac{r_1}{(j_1-r_1)\sqrt{(r_2-r_1)(r_3-r_1)}}= \frac{-j_1}{2j_1(\sqrt{2j_1})^2}=\frac{-1}{4j_1}. \end{equation*} Thus \begin{equation*}\begin{split} \phi_1=&\lim_{h\rightarrow0^+} \frac {-h}{4j_1\sqrt 8}\int_{r_1}^0 \frac{-1}{(j_1+\pi_3)\sqrt{\pi_3-r_1}}\,d\pi_3\\ =&\lim_{h\rightarrow0^+}\frac h{4j_1\sqrt 8} \frac 2{\sqrt{j_1+r_1}}\arctan \left.\left(\frac{\sqrt{\pi_3-r_1}}{\sqrt{j_1+r_1}}\right)\right|_{\pi_3=r_1} ^{\pi_3=0}\\ =&\frac h{4j_1\sqrt 8} \frac2{\sqrt{\frac{h^2}{32{j_1}^2}}}\frac\pi2\\ =&\frac\pi2. \end{split}\end{equation*} The right singularity of~\rf{557} arises corresponds to the singularity of the reduced spaces of the drift system, and it gives a phase shift of \begin{equation*}\begin{split} \phi_2&\equiv \lim_{h\rightarrow0^+}\int_0^{r_2}\frac{-h\pi_3}{({j_1}^2-{\pi_3}^2) \sqrt{8(\pi_3-r_1)(r_2-\pi_3)(r_3-\pi_3)}}\,d\pi_3\\ &=\frac{-h}{\sqrt8} \lim_{h\rightarrow0^+}\frac{\pi_3}{(j_1+r_2)\sqrt{r_2-r_1}} \lim_{h\rightarrow0^+}\int_0^{r_2}\frac{1}{(j_1-\pi_3) \sqrt{(r_2-\pi_3)(r_3-\pi_3)}}\,d\pi_3\\ &=\frac{-h}{\sqrt8}\frac{j_1}{2j_1\sqrt{2j_1}} \lim_{h\rightarrow0^+}\Biggl(\frac{-1}{\sqrt{(r_3-j_1)(j_1-r_2)}}\\ &\qquad\times\left.\arctan\left( \frac12\frac{\pi_3(2j_1-r_2-r_3)+2r_2r_3-j_1(r_2+r_3)} {\sqrt{(r_3-j_1)(j_1-r_2)(r_2-\pi_3)(r_3-\pi_3)}}\right) \right|_{\pi_3=0}^{\pi_3=r_2}\Biggr)\\ &=\frac{-h}{\sqrt8}\frac{j_1}{2j_1\sqrt{2j_1}} \frac{-1}{\frac{h}{4\sqrt{j_1}}}\frac{-\pi}{2}\\ &=-\frac\pi4, \end{split}\end{equation*} so that the phase shift as the homoclinic orbit is approached through positive $h$ is \begin{equation*} \phi=-2(\phi_1+\phi_2)=-2\left(\frac\pi2-\frac\pi4\right)=-\frac\pi2. \end{equation*} The same calculation for $h$ negative gives $\phi_1=-\pi/2$ and $\phi_2=\pi/4$ for a total phase shift of $\phi=\pi/2$. \footnotesize\frenchspacing
\section{Introduction} Cosmological $\gamma$-ray bursts (GRB) are believed to be produced in the fireballs of very energetic explosions, when a large amount of energy, $E\sim10^{51-54}\textrm{ erg}$, is released over a few seconds in a small volume in with a negligible baryonic load, $Mc^2\ll E$ (see Piran 1999 for a review). Most of the energy is eventually transferred to the baryons which are accelerated to ultra-relativistic velocities with a Lorentz factor $\gamma\simeq E/Mc^2\sim10^2$--$10^3$ (e.g., \cite{ShP90}; \cite{Paczynski90}). A substantial fraction of the kinetic energy of the baryons is transferred to a non-thermal population of relativistic electrons through Fermi acceleration at the shock (\cite{MR93}). The accelerated electrons cool via inverse Compton scattering and synchrotron emission in the post-shock magnetic fields and produce the radiation observed in GRBs and their afterglows (e.g., \cite{Katz94}; \cite{SNP96}; \cite{Vietri97}; \cite{Waxman97}; \cite{WRM97}). The shock could be either {\it internal} due to collisions between fireball shells caused by source variability (Paczy\'nski \& Xu 1994; \cite{RM94}), or {\it external} due to the interaction of the fireball with the surrounding interstellar medium (ISM; \cite{MR93}). The radiation from internal shocks can explain the spectra (Pilla \& Loeb 1998) and the fast irregular variability of GRBs (Sari \& Piran 1997a), while the synchrotron emission from the external shocks provides a successful model for the broken power-law spectra and smooth temporal behavior of afterglows (e.g., Waxman 1997a,b). In both cases, strong magnetic fields are required behind the shocks at all times in order to fit the observational data. The properties of the synchrotron emission from GRB shocks are determined by the magnetic field strength, $B$, and the electron energy distribution behind the shock. Both of these quantities are difficult to estimate from first principles, and so the following dimensionless parameters are often used to incorporate modeling uncertainties (\cite{SNP96}), \begin{equation} \epsilon_B\equiv\frac{U_B} {e_{\rm th}} ,\qquad \epsilon_e\equiv\frac{U_e}{e_{\rm th}} . \end{equation} Here $U_B={B^2}/{8\pi}$ and $U_e$ are the magnetic and electron energy densities and $e_{\rm th}=nm_pc^2(\bar\gamma_p-1)$ is the total thermal energy density behind the shock; where $m_p$ is the proton mass, $n$ is the proton number density, and $\bar\gamma_p$ is the mean thermal Lorentz factor of the protons. The observed afterglow spectra and lightcurves typically yield values of the magnetic energy parameter ranging from $\epsilon_B\sim0.1$ (Waxman 1997; Wijers \& Galama 1998), down to $10^{-2}$ (Granot et al. 1998) or even $\epsilon_B\sim10^{-5}$ (Galama et al. 1999; \cite{Vreeswijk99}) --- all below the {\em equipartition} limit $\epsilon_B\sim 1$. The existence of strong magnetic fields is naturally expected in the compact environments of potential GRB progenitors. First, the field might originate from a highly magnetized stellar remnant, such as a neutron star, with $B\la 10^{16}\textrm{ gauss}$. Second, a turbulent magnetic dynamo could amplify a relatively weak seed magnetic field in the vicinity of the progenitor. This process, however, requires the turbulence to be anisotropic and have a nonzero total helicity, ${\bf v\cdot(\nabla\times v)}\not=0$. A similar mechanism, called the $\alpha$-$\Omega$-dynamo, might operate in rapidly rotating objects (\cite{Thompson94}, see also \cite{MLR93}). Finally, the magnetic shearing instability (\cite{BalbusHawley91}) could amplify the magnetic field (but not the flux) in strongly sheared flows. The $e$-folding time for this instability is approximately the rotation period, which decreases with radius as $R^{-2}$ due to angular momentum conservation in the outflowing wind. Thus, being possibly important in the early stages of the fireball expansion (\cite{NPP92}), this instability is inefficient at large radii, where its $e$-folding time greatly exceeds the dynamical time of the fireball. In contrast to such progenitor environments where large magnetic fields are natural, there is currently no satisfactory explanation for the origin of the strong magnetic fields required in GRB shocks (see discussions in \cite{Thompson94}; \cite{MLR93}; \cite{SNP96}). Compression of the ISM magnetic field in external shocks yields a field amplitude $B\sim\gamma B_{\rm ISM}\sim 10^{-4} (\gamma/10^2) \textrm{ gauss}$, which is too weak (\cite{SNP96}) compared to the required equipartition value $B_{eq}\sim 50 (\gamma/10^2)(n_{\rm ISM}/1~{\rm cm^{-3}})^{1/2} \textrm{ gauss}$, and can account only for $\epsilon_B=(B/B_{eq})^2\la 10^{-11}$. Alternatively, some magnetic flux might originate at the GRB progenitor and be carried by the outflowing fireball plasma (or by a precursor wind). Because of flux freezing, the field amplitude would decrease as the wind expands. In this case, only a progenitor with a rather strong magnetic field $\sim 10^{16}\textrm{ gauss}$ might produce sufficiently strong fields during the GRB emission. However, since the field amplitude scales as $B\propto V^{-2/3}$ for an expanding shell of volume $V$, even a highly magnetized plasma at $R\sim 10^7\textrm{ cm}$ would possess only a negligible field amplitude of $\sim 10^{-2}\textrm{ gauss}$, or $\epsilon_{B}\la 10^{-7}$, at a radius of $R\ga 10^{16}\textrm{ cm}$, where the afterglow radiation is emitted\footnote{Both the magnetic field energy density and the thermal energy of the fireball scale as $\propto V^{-4/3}$ for adiabatic expansion. However, when shocks are generated, the plasma is heated due to the dissipation of the fireball kinetic energy, and the magnetic energy parameter decreases far below equipartition in the post shock gas.} (see also \cite{MLR93}). Moreover, the emitting material behind the external shock is continuously replenished by the ISM, and so the field originally carried by the fireball ejecta cannot account for the afterglow radiation. None of the above mechanisms is capable of generating near equipartition magnetic fields in the {\em external shocks} which produce the delayed afterglow emission. In this paper, we propose a different, {\em universal} mechanism of magnetic field generation in GRB shocks. It involves the relativistic generalization of the two-stream (Weibel) instability in a plasma. This instability is driven by the {\em anisotropy} of the Particle Distribution Function (PDF) and, hence, could operate in both internal and external shocks. Our main results are as follows: \begin{enumerate} \item The characteristic $e$-folding time in the shock frame for the instability is $\sim10^{-7}\textrm{ s}$ for internal shocks and $10^{-4}\textrm{ s}$ for external shocks. This time is much shorter than the dynamical time of GRB fireballs. \item The generated magnetic field is randomly oriented in space, but always lies in the plane of the shock front. \item The instability is powerful. It saturates only by nonlinear effects when the magnetic field amplitude approaches equipartition with the electrons (and possibly with the ions). \item The instability isotropizes the PDF and, thus, effectively heats the electrons and protons. \item The characteristic coherence scale of the generated magnetic field is of the order of the relativistic skin depth, i.e. $\sim10^3\textrm{ cm}$ for internal shocks and $\sim10^5\textrm{ cm}$ for external shocks. This scale is much smaller than the spatial scale of the source. \item The mean free path for Coulomb collisions is larger than the fireball size. However, the randomness of the generated magnetic field provides effective collisions due to pitch angle scattering of the particles in an otherwise collisionless plasma and, thus, justifies the use of the magneto-hydrodynamic (MHD) approximation for GRB shocks. The magnetic fields communicate the momentum and pressure of the outflowing fireball plasma to the ambient medium and define the shock boundary. \end{enumerate} The above mechanism results in tangential magnetic fields near the apparent limb of the source. Hence, the long-term synchrotron emission from the limb would be linearly polarized along the radial direction relative to the source center. Although the net polarization of a circularly symmetric source is zero, scattering of the radio afterglow emission of GRBs by the intervening Galactic interstellar medium would break the symmetry in the source image and result in polarization scintillations. This effect can be used to test the reality of our proposed mechanism for the generation of magnetic fields in GRB blast waves. The outline of the paper is as follows. The physical mechanism of the instability is discussed in \S \ref{S:WEIBEL}. The generation of magnetic fields in internal and external shocks is discussed in \S \ref{S:SHOCKS}. In \S \ref{S:PREDICT} we predict the polarization scintillation signal in our model. Finally, \S \ref{S:DISC} summarizes our main conclusions. \section{The Two-Stream Instability \label{S:WEIBEL}} The instability under consideration was first predicted by Weibel (1959) for a non-relativistic plasma with an anisotropic distribution function. The simple physical interpretation provided later by Fried (1959) treated the PDF anisotropy more generally as a two-stream configuration of a cold plasma. Below we give a brief, qualitative description of this two-stream magnetic instability. Let us consider, for simplicity, the dynamics of the electrons only, and assume that the protons are at rest and provide global charge neutrality. The electrons are assumed to move along the $x$-axis (as illustrated in Figure\ \ref{fig}) with a velocity ${\bf v}=\pm {\bf \hat x}v_x$ and equal particle fluxes in opposite directions along the $x$-axis (so that the net current is zero). Next, we add an infinitesimal magnetic field fluctuation, ${\bf B}= {\bf\hat z}B_z\cos(ky)$. The Lorentz force, $-e{{{\bf v}\over c}\times{\bf B}}$, deflects the electron trajectories as shown by the dashed lines in Figure\ \ref{fig}. As a result, the electrons moving to the right will concentrate in layer I, and those moving to the left -- in layer II. Thus, current sheaths form which appear to {\em increase} the initial magnetic field fluctuation. The growth rate is $\Gamma=\omega_{\rm p}v_y/c$, where $\omega_{\rm p}^2=(4\pi e^2n/m)$ is the non-relativistic plasma frequency (\cite{Fried59}). Similar considerations imply that perpendicular electron motions along $y$-axis, result in oppositely directed currents which suppress the instability. The particle motions along ${\bf\hat z}$ are insignificant as they are unaffected by the magnetic field. Thus, the instability is indeed driven by the PDF anisotropy and should quench for the isotropic case. The Lorentz force deflection of particle orbits increases as the magnetic field perturbation grows in amplitude. The amplified magnetic field is {\em random} in the plane perpendicular to the particle motion, since it is generated from a random seed field. Thus, the Lorentz deflections result in a pitch angle scattering which makes the PDF isotropic. If one starts from a strong anisotropy, so that the thermal spread is much smaller than the particle bulk velocity, the particles will eventually isotropize and the thermal energy associated with their random motions will be equal to their initial directed kinetic energy. This final state will bring the instability to saturation. We note the following points about the nature of the instability: \begin{enumerate} \item The instability is {\em aperiodic}, i.e., ${\rm Re}\,\omega=0$. Thus, it can be saturated by nonlinear effects only, and not by kinetic effects such as collisionless damping or resonance broadening. Hence, the magnetic field can be amplified up to high values. \item Despite its intrinsically kinetic nature, the instability is non-resonant,\footnote{This instability may be treated as an analog of the fire-hose instability in the absence of the external magnetic field.} i.e., it is impossible to single out a group of particles that is responsible for the instability. Since the bulk of the plasma participates in the process, the energy transferred to the magnetic field could be comparable to the total kinetic energy of the plasma. Hence, the instability is {\em powerful}. \item The instability is self-saturating. It continues until all the free energy due to the PDF anisotropy is transferred to the magnetic field energy. \item The generated magnetic field always lies in the plane perpendicular to the initial anisotropy axis of the PDF, i.e., to the shock propagation direction. \item The produced magnetic field is randomly oriented in the shock plane. The Lorentz forces randomizes particle motion over the pitch angle and, hence, introduces an effective scattering process into the otherwise collisionless system. This validates the use of the MHD approximation in the study of collisionless GRB shocks. \end{enumerate} \cite{SGbook} and \cite{MS63} provide a kinetic, non-relativistic treatment of the instability in both the linear and the quasi-linear regimes, and apply the theory of collisionless shocks to space plasmas. In the next section we will extend their analysis to the case of ultra-relativistic GRB shocks. \section{Magnetic Field Generation in GRB Shocks \label{S:SHOCKS}} We consider a GRB shock front expanding at a Lorentz factor, $\gamma_{\rm sh}$, behind which the particles have a thermal Lorentz factor, $\bar\gamma$. In this section, we will derive equations which are equally applicable to electrons and protons, whichever species dominates the growth of the instability. Later, we shall use the subscripts ``$e$'' and ``$p$'' to denote electrons and protons, respectively. We calculate all quantities in the comoving frame of the shock. A fully kinetic, relativistic treatment of the magnetic two-stream (Weibel) instability is a complicated task. The dispersion relation for a simplified ``water-bag'' PDF was derived by Yoon \& Davidson (1987) and is given by equation\ (\ref{disp}) in Appendix \ref{A1}. This dispersion relation implies that only a range of modes above a critical wavelength will grow [cf. Eq.\ (\ref{range})]. Naturally, the mode with the largest growth rate, $\Gamma_{\rm max}$, dominates and sets the characteristic length-scale of the magnetic field fluctuations, $\lambda\sim k_{\rm max}^{-1}$. The ultra-relativistic expressions for $\Gamma_{\rm max}$ and $k_{\rm max}$ are given by equation\ (\ref{gamma-k}) for a strong initial anisotropy. We write the corresponding $e$-folding time and correlation length of the field as \begin{equation} \tau\simeq\frac{\gamma_{\rm sh}^{1/2}}{\omega_{\rm p}} ,\quad \lambda\simeq2^{1/4}\frac{c\bar\gamma^{1/2}}{\omega_{\rm p}} . \label{scales} \end{equation} We can now estimate the nonlinear saturation amplitude of the magnetic field. The instability is due to the free streaming of particles. As the field amplitude grows, the transverse deflection of particles gets stronger, and their free streaming across the field lines is suppressed. The typical curvature scale for the deflections is the Larmor radius, $\rho=v_{\bot B}/\Omega_{\rm c}\simeq(\gamma_{\bot B}^2-1)^{1/2}mc^2/eB$, where $v_{\bot B}$ and $\gamma_{\bot B}$ are the transverse velocity and Lorentz factor of a particle relative to the local magnetic field. On scales larger than $\rho$, particles can only move along field lines. Hence, when the growing magnetic fields become such that $k_{\rm max}\rho\sim1$, the particles are magnetically trapped and can no longer amplify the field. Assuming an isotropic particle distribution at saturation ($\gamma_{\bot B}\sim\bar\gamma$), this condition can be re-written as \begin{equation} \frac{B^2/8\pi}{mc^2n(\bar\gamma-1)} \sim\frac{(\bar\gamma+1)}{2\sqrt2\,\bar\gamma} . \label{sat} \end{equation} For ${\bar\gamma}\gg 1$, this corresponds to a magnetic energy density close to equipartion with the amplifying particles. Interestingly, one may obtain the same result following a different analysis. First, the instability leads to a growth of the field amplitude [as given by the last term in Eq.\ (\ref{ke}), $\sim{\bf v\cdot}\partial_{\bf x} f$]. Second, nonlinearity leads to the transfer of energy to shorter wave lengths, $k>k_{\rm crit}$, where the fluctuations are damped [as described by the second term in Eq.\ (\ref{ke}), $\sim(e/c) {\bf v\times B\cdot}\partial_{\bf p} f$]. Thus, the steady value of $B$ is determined by balancing these two processes. Equating these two terms and replacing $\partial_x$ by $k_{\rm crit}\simeq k_{\rm max}/\sqrt2 \sim\rho_p/\sqrt{2}$, yields \begin{equation} v_{\bot B} k_{\rm crit}f\sim v_{\bot B}eBf/mc^2\gamma_{\bot B} . \end{equation} The field strength estimated here is equivalent to that given in equation\ (\ref{sat}) to within a factor of order unity. Direct computer simulations of the instability in both non-relativistic and relativistic electron plasmas confirm that the saturation occurs at slightly {\em sub-equipartition} values of $B$ (see, e.g., \cite{Califanoetal98}; Kazimura et al. 1998; \cite{Yangetal94}; \cite{WE91}), \begin{equation} \frac{B^2/8\pi}{mc^2n(\bar\gamma-1)}\equiv\eta\sim0.01 - 0.1\, . \label{field} \end{equation} where we introduced the efficiency factor $\eta\lesssim0.1$. The precise saturation level depends on the nonlinear modification of the PDF during the instability which is not accounted for by our linear analysis. We shall retain the efficiency factor, $\eta$, in our estimates. Note that the thermal Lorentz factor of particles, $\bar\gamma$, varies in time as the instability develops. Due to particle scattering by the generated magnetic fields, an initially highly anisotropic PDF with $\bar\gamma\ll\gamma_{\rm sh}$ will eventually evolve to an isotropic, ring-like distribution, for which $\bar\gamma\simeq\gamma_{\rm sh}$. Thus, the spatial scale and amplitude of the resultant magnetic field, given by equations\ (\ref{scales}) and (\ref{field}), will evolve during the lifetime of the instability because they are functions of $\bar\gamma$. In estimating these values at a GRB shock when the instability saturates, we take $\bar\gamma\simeq\gamma_{\rm sh}$. The $e$-folding time for the instability is independent of $\bar\gamma$ in the case of strong anisotropy. In the case of weak anisotropy, $\bar\gamma\approx\gamma_{\rm sh}$, the ``water-bag'' model used here is formally invalid, but the comparison of the ultra-relativistic results with non-relativistic results (e.g., Moiseev \& Sagdeev 1963) suggests that the instability quenches and the $e$-folding time scales as \begin{equation} \tau\simeq\lambda/c \propto\left[(\epsilon_\|-\epsilon_\perp)/\epsilon_\|\right]^{-3/2}, \end{equation} where $\epsilon_\|$ and $\epsilon_\perp$ are the average energies of particle motions along the direction of shock propagation and transverse to it. The field correlation length follows a similar scaling. The diffusive decay time of the generated magnetic field is $\tau_{\rm diff}\simeq1/\eta_Bk_{\rm max}^2$, where $\eta_B=mc^2\nu_{\rm coll}/4\pi ne^2$ is the magnetic diffusivity and $\nu_{\rm coll}$ is the particle collision frequency. Hence, the diffusion time \begin{equation} \tau_{\rm diff}\simeq\bar\gamma/\nu_{\rm coll} , \end{equation} is much longer than the fireball expansion time since the particle collision frequency in the fireball plasma is negligible. Thus, the magnetic field is not expected to dissipate its energy Ohmically over the fireball lifetime. Note that magnetic fields cannot be produced during the optically-thick phase of the fireball, because Compton scattering on the photons rapidly removes any anisotropy of the PDF. Next, we consider two types of GRB shocks in which magnetic fields might be generated. \subsection{Internal Shocks due to Shell Collisions Inside the Fireball} Rapid variability of a GRB source results in a fireball which is composed of thin layers (shells) moving with different Lorentz factors. To produce the observed non-thermal $\gamma$-ray spectrum, the shells must collide at sufficiently large radii where the internal shock region is optically-thin to both Compton scattering and $e^+e^-$-pair production. The collision should also occur before the fireball slows-down on the ambient medium. These conditions imply that the internal shock be mildly relativistic, with a Lorentz factor $\gamma_{\rm int}$ of order a few in the center of mass frame of the colliding shells (see Piran 1999 for more details). Prior to a collision, the electrons and protons in the colliding shells are cold relative to their bulk Lorentz factor, $\bar\gamma_{e,p}\lesssim\gamma_{\rm int}$. As typical parameters for the shells we assume a plasma density of $n\approx3\times10^{10}\textrm{ cm}^{-3}$, $\gamma_{\rm int}=4$, and initial thermal Lorentz factors $\bar\gamma_{p,e}\approx2$ (see e.g., Piran 1999; \cite{PL98}). The plasma frequencies for the electrons and protons are given by the relations, ${\omega_{\rm p}}_e=9.0\times10^3n^{1/2}\textrm{ s}^{-1}$ and ${\omega_{\rm p}}_p=2.1\times10^2n^{1/2}\textrm{ s}^{-1}$, where $n$ is in $\textrm{ cm}^{-3}$. For simplicity, we consider the collision of two identical shells. In the center of mass frame, the interaction of these collisionless shells yields a state of two inter-penetrating plasma streams, which is readily unstable to the generation of magnetic fields. Since ${\omega_{\rm p}}_e\gg{\omega_{\rm p}}_p$, the instability grows faster for the electrons than for the protons and so the electrons dominate the magnetic field generation process at early times. The electron instability saturates when the magnetic energy density becomes comparable to the electron energy density, $\gamma_{\rm int}n m_ec^2$. This energy is still much smaller than that associated with the protons\footnote{We assume that the dominant ion species in the relativistic GRB wind is protons. The generalization of our discussion to heavier ion species is straightforward.}. Thus, when the instability saturates for the electrons, it could still continue on a longer time-scale for the protons. The protons dominate energetically and could lead to near equipartition magnetic energy with \begin{equation} \epsilon_B\lesssim \eta\sim 0.1. \end{equation} From equation\ (\ref{scales}) we get the characteristic scale length and growth time of the instability for the protons, \begin{mathletters} \begin{eqnarray} \lambda&\simeq&2\times10^3 \left(\frac{n}{3\times10^{10}\textrm{ cm}^{-3}}\right)^{-1/2} \left(\frac{\gamma_{\rm int}}{4}\right)^{1/2}\textrm{ cm} ,~~~~~~~ \\ \tau&\simeq&6\times10^{-8} \left(\frac{n}{3\times10^{10}\textrm{ cm}^{-3}}\right)^{-1/2} \left(\frac{\gamma_{\rm int}}{4}\right)^{1/2}\textrm{ s}. \end{eqnarray} \end{mathletters} These quantities are decreased by a factor of $(m_p/m_e)^{1/2}=43$ for the electrons. The generation of magnetic fields in counter-streaming, electron-positron plasmas has been extensively studied numerically using particle-in-cell codes (e.g., \cite{Kazimuraetal98}; \cite{Califanoetal98}; \cite{Yangetal94}). A clear visual demonstration of the magnetic field amplification process is provided by Figure 2 of Kazimura et al. (1998). The rapid generation of a strong, small-scale magnetic field occurs at the interface of the colliding streams, and is followed by the gradual modification of the field structure around the interface, due to the nonlinear saturation and relaxation of the particle velocity anisotropy. The inferred value of $\eta\sim0.01-0.1$ is generic (\cite{Kazimuraetal98}; \cite{Yangetal94}). The amplication process produces also random electric fields with an energy density that is at most comparable to that of the magnetic component, $\langle E^2\rangle\sim(v/c)\langle B^2\rangle$ (\cite{Kazimuraetal98}). Unfortunately, no simulations were performed so far for colliding electron-proton plasmas. Numerical simulations of a plasma with species of somewhat different masses suggests that the energetics of the process is indeed dominated by the heavier species (\cite{Arons96}). Nevertheless, direct relativistic simulations with dynamical protons and electrons are required in order to assess the saturation amplitude of the magnetic field in GRB shocks of different properties. If the colliding shells do not possess similar densities, then the growth rate of the instability decreases or even shuts off beyond a particular density contrast, as discussed in Appendix \ref{A2}. In this regime, the shock may be dominated by electrostatic (Langmuir) turbulence. Unless the outflowing plasma is already contaminated by strong magnetic fields, the synchrotron emission from the collision of shells with very different densities would therefore be weak. \subsection{External Shock due to the Interaction of the Fireball with the ISM} Eventually, the fireball slows down due to its interaction with the surrounding ISM. The external shock produced by this interaction yields the delayed afterglow emission and is assumed to carry a strong magnetic field. As the shock propagates into the ISM, the fresh electrons and protons are reflected from the magnetized shock front back into the ISM. Thus, a two-stream state forms in the comoving frame of the shock and a magnetic field is amplified in the ISM, just in front of the shock. We assume that the fraction of reflected particles is of order unity, and so the above two-stream state is analogous to that produced in internal shocks\footnote{The existence of a shock discontinuity relies on the fact that the fraction of scattered particles at the shock front is close to unity. The relevant scattering process could be produced by either strong Langmuir or magnetic turbulence which mediates the pressure force of the post-shock gas to the pre-shock gas.}. The instability first acts on the electrons. The correlation scale and saturation amplitude of the field are given by equations\ (\ref{scales}) and (\ref{field}). Since the magnetic field is generated upstream and then transported downstream, we need to take account of the compression factor at the shock. Given the jump conditions for a relativistic shock, we have $\lambda=\lambda'/4\gamma_{\rm sh}$ (where ``prime'' denotes the parameter in front of the shock), while the ratio of the magnetic to thermal energy remains constant. For an ISM density $n_{\rm ISM}\approx1\textrm{ cm}^{-3}$, we therefore get the following parameters behind the shock, \begin{mathletters} \begin{eqnarray} {\epsilon_B}_e&=&\eta\,\left(\frac{m_e}{m_p}\right) \simeq5.5\times10^{-5}\eta_{.1} , \label{sat-e} \\ \lambda_e&=&3\times10^5 \left(\frac{\gamma_{\rm sh}}{10}\right)^{-1/2} \left(\frac{n_{\rm ISM}}{1\textrm{ cm}^{-3}}\right)^{-1/2} \textrm{ cm} ,~~~~~~~ \\ \tau_e&=&4\times10^{-4} \left(\frac{\gamma_{\rm sh}}{10}\right)^{1/2} \left(\frac{n_{\rm ISM}}{1\textrm{ cm}^{-3}}\right)^{-1/2} \textrm{ s}. \end{eqnarray} \end{mathletters} where $\eta_{.1}\equiv (\eta/0.1)$ and the subscript ``$e$'' denotes amplification of the magnetic field by the electrons only. The magnetic energy parameter is still normalized relative to the proton thermal energy. When the instability of the electrons saturates, further amplification by the protons may become important. The magnetic field is amplified in a thin layer in front of the shock, the width of which is of order the Larmor radius of the protons\footnotemark\ at the shock. \footnotetext{Regardless of whether the electrons or protons contribute to the instability, the width of the shock front is set by the heavier protons. The electrons follow the protons to ensure quasi-neutrality of the plasma (an electric field forms which keeps the electrons tied to the protons). The magnetic field amplification by the electrons occurs in a pre-shock region of width $\sim\rho_e\ll\rho_p$, and the electrons have sufficient time to amplify the magnetic field up to their saturation amplitude.} The time available for field amplification is, thus, roughly the crossing time, $t_{\rm amp}\sim\rho_p/c\sim\bar\gamma_p(\bar\gamma_e\gamma_{\rm sh}\eta)^{-1/2} (m_p/m_e)^{1/2}(B_{{\rm sat,}e}/B) \tau_p \sim 3(m_p/m_e)^{1/2} (B_{{\rm sat,}e}/B) \tau_p$, where $B_{{\rm sat,}e}$ denotes the field strength after saturation on the electrons [as given by Eq.~(\ref{sat-e})]. On the other hand, the growth of the field to a sub-equipartition amplitude with the protons would take at least $t_{\rm growth}= \tau_p\ln(B_{{\rm sat,}p}/B_{{\rm sat,}e}) \sim\tau_p\ln(m_p/m_e)^{1/2} \sim 3.8\tau_p$, i.e. comparable to the time available for the amplification of $B_{{\rm sat,}e}$ up to $B_{{\rm sat,}p}=(m_p/m_e)^{1/2}B_{{\rm sat,}e}$. However, since the growth of the field near saturation is slower than that during the linear stage of the process, the Weibel instability may not be able to build the field up to equipartition with the protons and yield $\epsilon_B\sim\eta$. Whether maximal amplification of the magnetic field can occur in this environment is uncertain\footnotemark\ and can be found only through detailed numerical simulations. \footnotetext{The uncertain saturation level might be affected by energy exchange between of the protons and the electrons and the excitation of competing modes. Initially, the electron Larmor radius is smaller than the proton Larmor radius. A slight charge separation results in a strong electric field, which maintains the quasi-neutrality of the moving plasma. The electric field keeps the electrons and protons at the same bulk velocity, but might also heat the electrons up to equipartition with the protons. Values of $\epsilon_e\sim 0.1$ are indeed indicated by afterglow data (but could result also from Fermi acceleration of the electrons at the shock front). The accelerated electrons might then amplify the magnetic field further. Otherwise, the so-called low-hybrid plasma waves are excited in collisionless shocks with magnetized electrons and unmagnetized protons. These waves are generated by the protons and have a typical growth rate $\Gamma_{LH}\sim(\Omega_p\Omega_e)^{1/2} \sim{\omega_{\rm p}}_p$ for $B\sim B_{{\rm sat,}e}$, i.e., comparable to the two-stream instability growth rate. Such waves may carry a significant amount of energy and also transfer it to the electrons via resonant interactions. In addition, Langmuir (electrostatic) turbulence might be generated via the interaction of the low-density beam (ISM) with the high-density shocked material (see Appendix \ref{A2}) and, thus, lower the efficiency $\eta$. } We thus conclude that the most robust prediction for the value of the magnetic field energy is $\epsilon_B\sim \eta(m_e/m_p)$, but somewhat higher values are also possible, so that \begin{equation} 5\times10^{-5}\eta_{.1} \lesssim \epsilon_B \la 0.1\eta_{.1} . \label{eq:limits} \end{equation} The predicted range of $\epsilon_B\sim10^{-1}$--$10^{-5}$ matches the results from modeling of recent afterglow data. Wijers \& Galama (1998) show that X-ray to radio spectrum of GRB970508 afterglow is consistent with the values of $\epsilon_B\sim0.07$, indicating the proton dominated regime of field generation. The field energy density for GRB990123 and GRB971214 is estimated to be $\epsilon_B\sim10^{-5}$ (\cite{Galama99}), and for GRP980703 $\epsilon_B\sim6\times10^{-5}$ (\cite{Vreeswijk99}), consistently with the electron dominated regime. \section{Polarization Scintillations \label{S:PREDICT} } In the previous section, we have found that the characteristic time it takes the magnetic field to grow up to equipartition values is orders of magnitude shorter than the dynamical time-scale of GRB shocks. Hence, the growth of the field does not have observational consequences. Similarly, the typical correlation length of the magnetic field is much smaller than the source size and cannot be resolved. Thus, conventional light-curve observations are unable to test the magnetic instability mechanism. However, polarization measurements might be more promising, as we show next. \subsection{General Considerations} Synchrotron radiation produced by relativistic electrons is known to be highly polarized, predominantly in the direction perpendicular to the local magnetic field (\cite{Ginzburg-book}; \cite{RybickiLightman}). It was shown in \S \ref{S:WEIBEL} that the generated magnetic field is randomly oriented in the plane of the shock front. The afterglow radiation emitted by any infinitesimal section of the GRB blast wave is relativistically beamed to within an opening angle $\theta_{\rm b}\sim\gamma_{\rm sh}^{-1}\ll 1$. Hence, an external observer sees a conical section of the fireball, as defined by this opening angle. In addition, the rapid deceleration of the fireball reduces its surface brightness as it expands. For a particular observed time, emission along the line-of-sight axis to the source center suffers from the shortest geometric time-delay, and hence originates at a larger radius and is dimmer than slightly off-axis emission. The source therefore appears as a narrow limb-brightened ring (Waxman 1997c; Sari 1998; Panaitescu \& Meszaros 1998; Granot et al. 1998a). The outer cut-off of the ring is set by the sharp decline in the relativistic beaming at angles greater than $\gamma_{\rm sh}^{-1}$. Interestingly, the shock surface appears to a distant observer as almost perfectly aligned along the line-of-sight at the edge of the ring. This effect results from relativistic aberration (Rybicky \& Lightman 1979, p. 110), i.e. the Lorentz transformation of angles from the shock frame (in which the normal to the shock surface is inclined at an angle $\gamma_{\rm sh}^{-1}$ relative to the line-of-sight) to the observer frame. Therefore, at the limb-brightened edge of the ring, the small-scale magnetic field is oriented tangentially on the sky. Consequently, the random magnetic field does not average-out but rather produces linear polarization which is oriented radially from the center at any point on the ring. The resulting synchrotron radiation obtains a degree of polarization of \begin{equation} \pi_{\rm syn}=\frac{p_e+1}{p_e+7/3}\simeq72\% \label{eq:pi_syn} \end{equation} for the typical value of the power-law index, $p_e=2.5$, of the electron energy distribution, $dN_e/d\gamma_e\propto\gamma_e^{-p_e}$, in GRB sources. The two-stream mechanism for the amplification of the magnetic field can be tested only if the source is resolved, since the {\em net} polarization of a circularly-symmetric image is zero\footnote{A net polarization signal might still result from an asymmetric source (e.g., due to a misaligned jet) or due to an inverse cascade of the magnetic field to large scales (Gruzinov \& Waxman 1998).}. There are two ways for resolving a compact GRB source: (i) scintillations of radio afterglows due to electron density irregularities in the ISM of the Milky Way galaxy (Goodman 1997); and (ii) gravitational microlensing due to an intervening star along the line-of-sight (Loeb \& Perna 1998). Since lensing occurs only rarely, we focus our discussion on the first method. Observations of interstellar scintillations probe angular scales of order a few micro-arcseconds ($\ \mu {\rm as}$), far below the VLBI resolution ($\sim300\ \mu {\rm as}$). The interstellar scintillations arise when fluctuations in the electron density randomly modulate the refractive index of the turbulent ISM. As a result of random focusing and diffraction of the electromagnetic wave, a point source produces a spatial pattern of random bright and dim spots --- the speckle pattern. The source brightness fluctuates as the observer moves across the pattern. The characteristic angular correlation length of the pattern, $\theta_0$, is set by the statistical properties of the ISM turbulence. If, however, the source is extended, then the overall pattern is obtained from the superposition of the incoherent patterns of its individual parts. Thus, if the angular size of the source, $\theta_{\rm s}$, is larger than the characteristic scale of the speckles, namely $\theta_{\rm s}>\theta_0$, then the intensity fluctuations wash-out and the scintillation amplitude diminishes. The observations of a late-time decline in the amplitude of intensity scintillations for the radio afterglows GRB970508 (\cite{Frail-etal97}; \cite{Waxman-etal98}) and GRB980329 (\cite{Taylor-etal98}) provide an estimate for the shock radius, $R_{\rm s}\sim10^{17}\textrm{ cm}$ at times of $\sim1$ month and $\sim2$ weeks after these bursts, respectively. These estimates are consistent with the simplest fireball model predictions. The Weibel instability mechanism predicts that different segments of the ring-like source emit synchrotron radiation which is linearly polarized along the radial axis, so that the net polarization vanishes when averaged over the source. If $\theta_{\rm s}\ll\theta_0$, the source is effectively point-like and hence symmetric. This regime is characterized by strong intensity scintillations and weak polarization fluctuations. In contrast, when $\theta_{\rm s}>\theta_0$, different parts of the source are mapped differently, and the source is resolved. As the earth moves through the scintillation pattern, an observer will measure fluctuations in the direction and amplitude of the polarization, while the intensity would vary only weakly due to the overlap of the separate speckle patterns. The polarization scintillations should therefore be strong when the flux fluctuations are weak. We consider two types of scintillations, diffractive and refractive\footnote{Effects due to differential Faraday rotation or anisotropy of the ISM turbulence are unimportant because of the smallness of the scattering angle, $\sim \ \mu {\rm as}$ (\cite{Narayan99}).} (\cite{GN85}; \cite{BN85}). Diffractive scintillations occur when the source is nearly point-like, $\theta_{\rm s}\ll\theta_{\rm d}$, relative to \begin{equation} \theta_{\rm d}\simeq 3\left(\frac{\nu}{10\textrm{ GHz}}\right)^{-11/5} \ \mu {\rm as} , \end{equation} which is the diffraction angle for a typical scattering measure of $10^{-3.5} {\rm m}^{-20/3}~{\rm kpc}$ (\cite{Goodman97}). The flux modulation amplitude in the strong scattering regime is close to $100\%$. For a Kolmogorov spectrum of ISM turbulence, the characteristic speckle length is \begin{equation} \theta_0\simeq2.3\left(\frac{\nu}{10\textrm{GHz}}\right)^{6/5}\ \ \mu {\rm as}, \label{theta0} \end{equation} assuming a scattering screen distance of $\sim 1~{\rm kpc}$ and a typical scattering measure of $10^{-3.5} {\rm m}^{-20/3}~{\rm kpc}$ (Goodman 1997). The time-scale for diffractive scintillations is \begin{equation} t_{\rm diff}\simeq3\left(\frac{\nu}{10\textrm{ GHz}}\right)^{6/5}\textrm{ hr} \end{equation} if the transverse velocity of the line of sight is dominated by the earth with $v_\bot\simeq30\textrm{ km s}^{-1}$. As long as $\theta_{\rm s}\ll\theta_0$, the polarization is close to zero, but when the source approaches the speckle correlation length, $\theta_{\rm s}\sim\theta_0$, the polarization scintillations could grow up to a large amplitude, of order a few tens of percents [cf. Eq.~(\ref{eq:pi_syn})]. For these scintillations to be detected, the source must be observed at relatively low frequencies (\cite{Goodman97}), namely $\nu\lesssim10\textrm{GHz}$ for typical ISM conditions. Unfortunately, the synchrotron self-absorption often occurs at frequencies below 5~GHz, and so the afterglow might be fainter at these low frequencies, making the detection of polarization scintillations more difficult. In addition, the source image resembles more a filled disk rather than a hollow ring at low frequencies (Granot et al. 1998a). The unpolarized radiation emitted near the center of the disk will thus lower the overall degree of polarization. As the source gets larger, $\theta_{\rm s}\gg\theta_{\rm d}$, the diffractive effect weakens, and the scintillations are dominated by the refractive effect, which yields only modest intensity fluctuations with an amplitude $\sim10\%$. The polarization fluctuations in this regime have a corresponding amplitude of only a few percents. The characteristic time-scale for the refractive modulation is \begin{equation} t_{\rm ref}\simeq14\left(\frac{\theta_{\rm eff}}{10\ \mu {\rm as}}\right)\textrm{ hr}, \end{equation} where $\theta_{\rm eff}$ is the effective size of the source (see \cite{Goodman97} for details). \subsection{Polarization Scintillations: Formalism} The properties of the radiation field are fully described by four {\em scalar} parameters --- the Stokes parameters (\cite{Ginzburg-book}), which are {\em additive} for incoherent sources. For synchrotron radiation produced by relativistic electrons, these parameters include the intensity $I$ and \begin{mathletters} \begin{eqnarray} Q&=&I\,\cos 2\psi\,\cos 2\chi,\\ U&=&I\,\cos 2\psi\,\sin 2\chi,\\ V&=&0. \end{eqnarray} \label{Stokes} \end{mathletters} The last parameter, $V$, describes circular polarization while $Q$ and $U$ describe linear polarization. $\chi$ is the angle between the polarization axis and an arbitrary fixed direction in the sky and $\cos 2\psi=(I_{\|}-I_{\perp})/(I_{\|}+I_{\perp})$ is the difference between the radiation intensity along the two orthogonal axes of polarization divided by the sum (see \cite{Ginzburg-book}; and \cite{RybickiLightman}, p. 180). Both $\chi({\bf r})$ and $\psi({\bf r})$ are determined by the source, but are not affected by the scintillations. The degree of polarization is defined as \begin{equation} \pi=\left(Q^2+U^2+V^2\right)^{1/2}/I . \label{degree-def} \end{equation} Given a power spectrum of electron density fluctuations in the ISM, the statistics of speckles in a scintillation pattern is usually characterized by the second moment correlation of the complex electric field of the electromagnetic radiation, \begin{eqnarray} {\sf W}(\Delta{\bf x})&=&{\overline{E({\bf x})E^*({\bf x}+\Delta{\bf x})}} \nonumber\\ &\propto&\exp\left[-D_\varphi(\Delta{\bf x})/2\right] \nonumber\\ &\propto&\exp\left[-const\times (|\Delta{\bf x}|)^{\beta-2}\right], \label{eq:beta} \end{eqnarray} where ${\bf x}$ and $\Delta{\bf x}$ are two-dimensional vectors on the plane normal to the line of sight, the ``bar'' denotes an ensemble average, and $\beta$ is the power-law index of the power spectrum of electron density fluctuations, $|\delta n_e(q)|^2\propto q^{\beta}$ with $q$ being the spatial wave-number. The quantity $D_\varphi$ is the {\em phase structure function} which yields the phase shift along different paths and is determined by the ISM turbulence. The inferred value of $\beta$ for the Galactic ISM is somewhat uncertain but close to the Kolmogorov theory prediction $\beta=11/3$ (\cite{Armstrong-etal95}). In calculating the scintillation indexes below, we adopt the approximate value of $\beta\approx 4$ for which the ${\sf W}$ is Gaussian, which greatly simplifies the calculation. The Fourier transform of ${\sf W}$ is the apparent brightness distribution of the scattered image of a point source: \begin{equation} W(\theta,\phi)=\left(I_0-\overline{I_0}\,\right)/\,\overline{I_0} \rightleftharpoons{\sf W}, \end{equation} where $\rightleftharpoons$ denotes a Fourier conjugated pair, and $\theta=r/const$ and $\phi$ are the radial and angular polar coordinates on the sky relative to the source center. The scattered image of an extended source is the convolution of the image kernel of a point source with the brightness distribution at the source, $P_I(\theta,\phi)$, \begin{mathletters} \begin{eqnarray} I(\theta,\phi)&=&W(\theta,\phi)\ast P_I(\theta,\phi) \nonumber\\ &\equiv&\int\!\!\! \int W(\theta-\theta',\phi-\phi')\,P_I(\theta',\phi')\, \theta'{\rm d} \theta'\, {\rm d} \phi' . \nonumber\\& & \end{eqnarray} Similarly, the ``images'' of the other Stokes parameters are \begin{equation} Q(\theta,\phi)=W\ast P_Q ,\qquad U(\theta,\phi)=W\ast P_U . \end{equation} \label{IQU} \end{mathletters} Finally, the amplitude of the intensity fluctuations due to scintillations is determined by the so-called {\em scintillation index}: \begin{equation} S_I=\left({\langle I^2\rangle\over \langle W^2\rangle}\right)^{1/2} \label{Si} \end{equation} with analogous definitions for the indexes of the other Stokes parameters $S_Q$ and $S_U$. We use angular brackets to denote integrals of the form, $\langle W^2\rangle\equiv\int[W(\theta,\phi)]^2\,\theta{\rm d}\theta \,{\rm d}\phi$. The normalized amplitude of the polarization scintillations is described by the scintillation indexes of the polarization signal $S_{QU}$ and the degree of polarization $S_\pi$, \begin{equation} S_{QU}\equiv\left(S_Q^2+S_U^2\right)^{1/2} , \qquad S_\pi=S_{QU}/S_I . \label{Squ} \end{equation} \subsubsection{Polarization Scintillations of GRB Afterglows} To illustrate the qualitative properties of the polarization scintillations in GRB afterglows we consider a crude model for the source that simplifies the related integrals considerably. We approximate the circular source as having a uniform surface brightness over the region $0<\theta<\theta_{\rm s}(t)$, on the sky. We also normalize the total flux to unity at all times since it enters only as a multiplicative factor to the polarization indexes. The linear polarization is oriented along the radial direction, so that the polarization angle is equal to the polar angle $\chi\equiv\phi$ in equations\ (\ref{Stokes}), and the degree of polarization is assumed to be constant over the source, $\pi_{\rm s}=0.72$ [cf. Eq.~(\ref{eq:pi_syn})]. Much of the radiation from the ring-like image of a real source acquires this polarization level, although the overall polarization is somewhat degraded by emission from the central part of the ring. Our estimates should therefore be regarded as an upper limit on the measurable polarization amplitude. The brightness distribution function for the scattered image of a point source, $W$, is taken to be a Gaussian with a variance set by the speckle angular scale, $W=\exp[-\theta^2/\theta_0^2]$. The angular size of the source as a function of time, $\theta_{\rm s}(t)$, was evaluated by Waxman et al. (1998). For a cosmological source at a redshift $z_{\rm s}\sim1$, it reads \begin{equation} \theta_{\rm s}\simeq1.4\left(\frac{E}{10^{52} \textrm{ erg}}\right)^{1/8} \left(\frac{n_{\rm ISM}}{1 \textrm{ cm}^{-3}}\right)^{-1/8} \left({t\over 1~{\rm week}}\right)^{5/8} \ \mu {\rm as}, \end{equation} where $E$ is the total energy of the fireball and $t$ is elapsed time from the detection of the explosion. The scintillation indexes can then be numerically calculated as functions of $\theta_{\rm s}(t)/\theta_0$, using equations~(\ref{Stokes})--(\ref{Squ}). The temporal evolution of the scintillation indexes for a source with $z_{\rm s}=1$, $E=10^{52}~{\rm ergs}$ and $n_{\rm ISM}=1~{\rm cm^{-3}}$ is presented in Figure\ \ref{scint}. At early times, when the source size is small ($\theta_{\rm s}\ll\theta_0$), the polarization fluctuations are weak while the intensity fluctuations are at maximum. When the source size approaches the diffractive scattering angle, $\theta_d$, the source is resolved and the observed radiation is partially polarized. At the same time, the intensity fluctuation amplitude declines due to the overlap between speckles. The polarization fluctuations peak when $\theta_{\rm s}\sim\theta_0$ at a value of $\sim20\%\times (\pi_{\rm s}/0.72)$. As the source size increases even further, the fluctuation amplitude of both the intensity ($S_{I}$) and the polarization ($S_{QU}$) decrease, due to the overlap of scattering patterns from different regions of the source. However, the fluctuation level of the {\em degree} of polarization ($S_\pi$) continues to increase with increasing source size and asymptotes at $\sim \pi_{\rm s}=72\%$. Thus, the saturation level of $S_\pi$ is independent of the details of the scattering processes and provides information about the intrinsic degree of polarization at the source. \section{Conclusions \label{S:DISC}} We have shown that the relativistic two-stream magnetic instability is capable of producing strong magnetic fields in the internal and external shocks of GRB sources. The generated fields are randomly oriented in the plane of the collisionless shock front, and fluctuate on scales much smaller than the size of the emission region. The instability inevitably produces magnetic fields with the magnetic energy parameter of $\epsilon_B \sim 10^{-5}$--$10^{-4}$ due to the isotropization of the electrons at the shock (see, e.g., the simulations by Kazimura et al. 1998), and could saturate at yet higher values of $\epsilon_B\la 0.1$ if the protons do the same. Numerical simulation of electron-proton plasmas are necessary in order to examine under which conditions the protons might enhance the magnetic energy up to these high values. Galama et al. (1999) suggested a distinction between two classes of GRB afterglows: radio-weak GRBs like GRB971214 or GRB990123 where the magnetic energy parameter might be as low as $\epsilon_B\sim 10^{-6}$--$10^{-5}$, and radio-loud GRBs like GRB970508 where $\epsilon_B \sim 10^{-1}$ (Waxman 1997a,b; Wijers \& Galama 1998; Sari et al. 1998). Low-field afterglows are short and dim in the radio (and account for the majority of the afterglow population) while high-field afterglows are long-lived and bright in the radio. In our model, low-field GRBs would arise naturally due to the saturation of the instability at the initial kinetic energy of the electrons. High-field afterglows might result from proton amplification of the magnetic energy. Our model for the magnetic field generation predicts the existence of polarization scintillations in the radio afterglows of GRBs. Since the typical correlation length of the generated magnetic field is very small, no net polarization is expected in the absence of scintillations, unless the circular symmetry of the source is broken (e.g. due to a jet which is misaligned with the line-of-sight) or if there is an inverse cascade of the generated magnetic field to much larger scales. In the absence of such complications, the polarization scintillations should appear typically after a week, when the angular size of the source becomes of order a micro-arcsecond, or equivalently when its physical size is $\sim10^{17}\textrm{ cm}$. The normalized amplitude of the polarization scintillation signal at that time could be as high as $\sim 10$--$20\%$. \acknowledgements We thank Ramesh Narayan, Martin Rees, and Pawan Kumar for insightful comments, and Dale Frail, Bohdan Paczy\'nski, Eli Waxman, Ralf Wijers, Valentin Shevchenko, and Vitaly Shapiro for useful discussions. This work was supported in part by NASA ATP grants NAG 5-7768 and NAG 5-7039 (for AL) and NAG 5-3516 (for MM). \begin{appendix} \section{Ultra-relativistic Treatment of the Magnetic Instability \label{A1}} Starting with the kinetic equation \begin{equation} \partial_t f + {\bf v\cdot}\partial_{\bf x} f + (e/c){\bf v\times B\cdot}\partial_{\bf p} f = 0, \label{ke} \end{equation} for the collisionless plasma, separating the PDF into an unperturbed part and an infinitesimal perturbation, $f=F({\bf p})+\tilde f$, and specifying $F({\bf p})$, one can obtain (\cite{YD87}) the following dispersion relation for the magnetic (Weibel) instability in the relativistic regime: \begin{equation} 1=\frac{c^2k^2}{\omega^2}+\frac{\omega_{\rm p}^2/\hat\gamma}{\omega^2} \left(G(\beta_\bot)+\frac{1}{2}\frac{\beta_\|^2}{(1-\beta_\bot^2)} \left[\frac{c^2k^2-\omega^2}{\omega^2-c^2k^2\beta_\bot^2}\right]\right) , \label{disp} \end{equation} where $ \beta_\|=p_\|/\hat\gamma mc, \ \beta_\bot=p_\bot/\hat\gamma mc, \ \hat\gamma=(1-\beta_\|^2-\beta_\bot^2)^{-1/2}, \ G(\beta_\bot)=(2\beta_\bot)^{-1}\ln\! \left[(1+\beta_\bot)/(1-\beta_\bot)\right]$, and $p_\|$ abd $p_\bot$ are the components of particle momentum averaged over the PDF. Here we denote quantities parallel and perpendicular with respect to the direction of the shock propagation, opposite to the convention used by \cite{YD87}. It is easy to demonstrate that the instability occurs for the range of $k^2$ given by \begin{equation} 0<k^2<k^2_{\rm crit}\equiv \left(\frac{\omega_{\rm p}^2}{\hat\gamma c^2}\right) \left[\frac{\beta_\|^2}{2\beta_\bot^2(1-\beta_\bot^2)}-G(\beta_\bot)\right], \label{range} \end{equation} and only with anisotropic PDFs for which the expression in square brackets is positive. The mode with the largest growth rate dominates in the evolution. We therefore want to find the maximum growth rate, $\Gamma_{\rm max}$, and the corresponding wave vector of the fastest growing mode, $k_{\rm max}$. Upon straightforward but lengthy calculations, we obtain: \begin{mathletters} \begin{eqnarray} \Gamma_{\rm max}^2&=&\frac{\omega_{\rm p}^2}{\hat\gamma(1-\beta_\bot^2)} \left[\frac{\beta_\|^2}{1-\beta_\bot^2}+2\beta_\bot^2G(\beta_\bot)- \frac{2\sqrt{2}\beta_\|\beta_\bot}{(1-\beta_\bot^2)^{3/2}} \left(\frac{\beta_\|^2\beta_\bot^2}{1-\beta_\bot^2}+ \left(1-2\beta_\bot^2-\beta_\bot^4\right)G(\beta_\bot)\right)^{1/2}\right], \nonumber\\ \\ k_{\rm max}^2&=&\frac{\omega_{\rm p}^2}{\hat\gamma c^2(1-\beta_\bot^2)} \left[\frac{-\beta_\|^2}{2(1-\beta_\bot^2)}-G(\beta_\bot)+ \frac{(1+\beta_\bot^2)\beta_\|}{\sqrt{2}(1-\beta_\bot^2)^{3/2}} \left(\frac{\beta_\|^2}{1-\beta_\bot^2}+ \frac{1-2\beta_\bot^2-\beta_\bot^4}{\beta_\bot^2}\: G(\beta_\bot)\right)^{1/2}\right] . \nonumber\\ \end{eqnarray} \end{mathletters} These exact equations may be greatly simplified by assuming that the plasma is ultra-relativistic and the particle parallel momenta (associated with the bulk motion) are much larger than their perpendicular ones (due to their thermal motion): $\gamma_\|\gg\gamma_\bot\gg1$. Then $\hat\gamma\simeq\gamma_\|=\gamma$, and we readily obtain, \begin{equation} \Gamma_{\rm max}^2\simeq\frac{\omega_{\rm p}^2}{\gamma} \left(1-2\sqrt{2}\frac{\gamma_\bot}{\gamma}\right) , \qquad k_{\rm max}^2\simeq\frac{1}{\sqrt{2}}\frac{\omega_{\rm p}^2}{\gamma_\bot c^2} \left(1-\frac{3}{\sqrt{2}}\frac{\gamma_\bot}{\gamma}\right) . \label{gamma-k} \end{equation} Note that in the second equation, $\omega_{\rm p}^2$ is divided by $\gamma_\bot$, which is much smaller than $\gamma$. \section{Asymmetric Two-stream Instability \label{A2}} \subsection{Cold beam -- Plasma Instability } Here we consider the case when two interpenetrating collisionless plasma streams have different densities and speeds in the center of mass frame. Instabilities which occur in such a situation are often referred to as beam -- plasma instabilities. The lack of symmetry in the system complicates analytical, fully relativistic analysis and requires numerical simulations. Below we provide quantitative estimates based on extrapolation of the nonrelativistic results to the ultra-relativistic case. The non-relativistic case of a beam--plasma instability has been considered in different regimes (see e.g., \cite{Akhiezer-book}). If the densities of the two streams are very different from each other, the center of mass frame coincides with the rest frame of the denser stream, which we refer to as the ``bulk plasma.'' The lower density stream is moving with some velocity $u$ relative to it and is referred to as ``beam''. We denote the parameters of the beam by a prime. The dispersion relation for the magnetic instability in the case of a cold beam reads (\cite{Akhiezer-book}, v.1, p.306) \begin{equation} \omega^2=-{\omega'^2_{\rm p}}_e\left(\frac{k^2u^2} {k^2c^2+{\omega'^2_{\rm p}}_e} -\frac{k^2{v_{\rm th}^2}_e{v_{\rm th}^2}_p} {{\omega^2_{\rm p}}_e{v_{\rm th}^2}_p + {\omega^2_{\rm p}}_p{v_{\rm th}^2}_e} \right), \end{equation} where $u$ is the beam velocity. We can then find the maximal growth rate and the fastest growing mode, as in Appendix \ref{A1}, \begin{equation} \Gamma_{\rm max}^2=k^2_{\rm max}c^2 \simeq{\omega_{\rm p}}_e{\omega'_{\rm p}}_e(u/{v_{\rm th}}_e) . \end{equation} This result suggests the following scalings with the density ratio of the beams \begin{equation} \Gamma_{\rm max}\propto k_{\rm max}\propto \left(n_e'/n_e\right)^{1/4}, \qquad \epsilon_B\propto \left(n_e'/n_e\right)^{1/2}. \label{re-scale} \end{equation} \subsection{Hot beam -- Plasma Instability } When particle pitch-angle scattering at a shock is strong, the beam becomes ``hot'', $u\sim v'_{{\rm th}_e}\gg v_{{\rm th}_e}$. Then, the dispersion relation becomes (\cite{Akhiezer-book}) \begin{equation} \omega=i\,\sqrt{\frac{2}{\pi}}\, \frac{k{v'^2_{\rm th}}_e}{{\omega'^2_{\rm p}}_e({v'^2_{\rm th}}_e+u^2)} \left(\frac{u^2}{{v'^2_{\rm th}}_e}\,{\omega'^2_{\rm p}}_e-k^2c^2 -{\omega^2_{\rm p}}_e\right), \qquad\textrm{where } \quad k^2c^2+{\omega^2_{\rm p}}_e\approx \frac{u^2}{{v'^2_{\rm th}}_e}\,{\omega'^2_{\rm p}}_e . \end{equation} The instability occurs when \begin{equation} k^2c^2+{\omega^2_{\rm p}}_e <\frac{u^2}{{v'^2_{\rm th}}_e}\,{\omega'^2_{\rm p}}_e . \end{equation} Thus, the instability {\em shuts off} for $k\to0$ when \begin{equation} {\omega'_{\rm p}}_e/{\omega_{\rm p}}_e =(n'/n)^{1/2}< {{v'_{\rm th}}_e}/u\lesssim1, \end{equation} which is satisfied when $n'/n\lesssim1$. In this case, however, Langmuir (longitudinal, electrostatic, high-frequency) waves are efficiently generated with the (maximum) growth rate comparable to that of magnetic instability in the previous cases: \begin{equation} \Gamma_{\rm Langmuir}\simeq\frac{3^{1/2}}{2^{4/3}}\left(\frac{n'_e}{n_e}\right)^{1/3} {\omega_{\rm p}}_e . \end{equation} Random electric fields of Langmuir turbulence scatter plasma particles and provide effective collisions at the shock, so that the MHD approximation is applicable. A detailed analysis of this process is, however, beyond the scope of this paper. \end{appendix}
\section{Introduction} The problem of the quantum state of a nucleating bubble has been addressed in the literature several times\cite{coleman,rubakov,VV,MT,TMdecay}. The results relevant for our discussion can be summarized as follows. We have a self-interacting scalar field $\sigma$ (the tunneling field) described by the lagrangian \begin{equation} {\mathcal L}_\sigma = -\frac{1}{2}\partial_\mu\sigma\partial^\mu\sigma-V(\sigma) \end{equation} where $V(\sigma)$ has a local (metastable) minimum at some value $\sigma_F$ and a global one at $\sigma_T$ (see Fig.~\ref{potencial}). The bubble nucleation can be pictured as the evolution of the $\sigma$ field in imaginary time. The solution of the corresponding Euclidean time equation which interpolates between the false vacuum at spacetime infinity and the true vacuum inside the bubble is called the bounce. In the absence of gravity, vacuum decay is dominated by the $O(4)$ symmetric bounce solution \cite{colemanmin}. So we shall write the tunneling field as a function of $\tau\equiv(T_E^2+{\bf X})^{1/2}$ alone, \begin{equation} \sigma = \sigma_0(\tau), \end{equation} where $(T_E,{\bf X})$ are Cartesian coordinates in Euclidean space. The solution describing the bubble after nucleation is given by the analytic continuation of the bounce to Minkowski time $T$ through the substitution $T_E=-iT$. Then, the bubble solution depends only on the Lorentz invariant quantity $({\bf X}^2-T^2)^{1/2}$, where $(T,{\bf X})$ are the usual Minkowski coordinates. If there are quantum fields interacting with the tunneling field, their state will be significantly affected by the change of vacuum state. Pioneering investigations of this matter were carried out by Rubakov \cite{rubakov} and Vachaspati and Vilenkin \cite{VV}. These latter authors considered a model of two interacting scalar fields $\sigma$ and $\Phi$, and found the quantum state for $\widehat{\bf \Phi}$ (the quantum counterpart of $\Phi$) by solving its functional Scr{\"o}dinger equation. In order to find a solution, they impose as boundary conditions for the wave function $\Psi(\tau;\Phi]$ regularity under the barrier and the tunneling boundary condition (see \cite{VV} for details). They found that the quantum state must be SO(3,1) invariant. \begin{figure}[t] \centering \leavevmode\epsfysize=5cm \epsfbox{potencial.eps}\\[3mm] \begin{quote} {\sl \caption[fig1]{\label{potencial} Assumed shape for the potential of the tunneling field. It has a local minimum which corresponds to the false vacuum at $\sigma_F$ and a global minimum, the true vacuum, at $\sigma_T$. The bounce corresponds to the Euclidean evolution of the tunneling field under the barrier.}} \end{quote} \end{figure} A somewhat different approach was pursued later by Sasaki and Tanaka \cite{MT}. They carried out a refinement of the method for constructing the WKB wave function for multidimensional systems, first introduced by Banks, Bender and Wu \cite{banks} and extended to field theory by Vega, Gervais and Sakita \cite{vega}, and obtained the so called quasi-ground state wave function. The quasi-ground state wave function is a solution of the time independent functional Schr{\"o}dinger equation to the second order in the WKB approximation which is sufficiently localized at the false vacuum so that it would be the ground state wave functional it there were no tunneling. They also found that the state must be SO(3,1) invariant. Moreover, general arguments, due to Coleman \cite{gencol}, suggest that the decay must be SO(3,1) invariant. If not, the infinite volume Lorentz group will make the nucleation probability diverge. From a practical point of view, therefore, it would be interesting to know to what extent symmetry considerations alone can be used to determine the quantum state after nucleations. As a first approach to this question it will be useful to compute the two-point function and the renormalized expectation value of the stress tensor in a SO(3,1) invariant quantum state for two simples models of one-bubble spacetimes. \section{General Formalism} Our aim is to study the quantum state of a field $\Phi$ described by a Lagrangian of the general form \begin{equation} {\mathcal L}_\Phi=-\frac{1}{2}\partial_\mu\Phi\partial^\mu\Phi - \frac{1}{2}m(\sigma)^2 \Phi^2,\label{wish} \end{equation} where the mass term is due to the interaction of the field $\Phi$ with a nucleating bubble. Working from the very beginning in the Heisenberg picture, we will construct an SO(3,1) invariant quantum state for the field $\widehat{\bf \Phi}$. After we will find its Hadamard two-point function $G^{(1)}(x,x')\equiv \langle 0|\{\widehat{\bf\Phi},(x)\widehat{\bf\Phi}(x')\}|0\rangle$, and we will check whether it is of the Hadamard form \cite{hadamardA,hadamardB,hadamardC,waldkay}. Loosely speaking, a Hadamard state can be described\footnote{For a more precise definition of Hadamard states see \cite{waldkay}.} as a state for which the singular part of $G^{(1)}(x,x')$ takes the form \begin{equation} G^{(1)}_{\rm sing}(x,x') = \frac{u}{\sigma}+v\log(\sigma), \end{equation} where $\sigma$ denotes half of the square of the geodesic distance between $x$ and $x'$, and $u$ and $v$ are smooth functions that can be expanded as a power series in $\sigma$, at least for $x'$ in a small neighborhood of $x$. Hadamard states are considered physically acceptable because for them the point-splitting prescription gives a satisfactory definition of the expectation value of the stress-energy tensor. After clarifying the singular structure of $G^{(1)}(x,x')$, we will use the point-splitting formalism \cite{christensen,waldreg,folacci,dorca} to compute the renormalized expectation value of the energy-momentum tensor in this quantum state. Finally we will briefly discuss the applicability of a uniqueness theorem for quantum states due to Kay and Wald \cite{waldkay}. \section{SO(3,1) coordinates} In the present paper we will restrict ourselves to piecewise flat spacetime. It proves very useful to use coordinates adapted to the symmetry of the problem. So we will coordinatize flat Minkowski space using hyperbolic slices, which will embody the symmetry under Lorentz transformations. We define the new coordinates $(t,r)$ (Milne coordinates) by the equations \begin{equation} t \equiv (T^2-{\bf X}^2)^{1/2} \hskip 1cm r\equiv \mbox{tanh}^{-1} (|{\bf X}|/T), \end{equation} where $(T,{\bf X})$ are the usual Minkowski coordinates. In terms of these coordinates, we have \begin{equation} ds^2=-dt^2+t^2d\Omega_{H_3}, \end{equation} where \begin{equation} d\Omega_{H_3}=dr^2+\sinh^2r\,d\Omega_{S_2} \end{equation} is the metric on the unit 3-dimensional spacelike hyperboloid, and $d\Omega_{S_2}$ is the line element on a unit sphere. The above coordinates cover only the interior of the lightcone from the origin. In order to cover the exterior, we will use the Rindler coordinates \begin{equation} \xi_R\equiv({\bf X}-T^2)^{1/2} \hskip 1cm \chi_R \equiv \mbox{tanh}^{-1}(T/|{\bf X}|). \end{equation} In terms of this coordinates, the line element reads \begin{equation} ds^2 = d\xi_R^2 + g_{AB} dx^A dx^B = d\xi_R^2 + \xi_R^2\,d\Omega_{dS_3}, \end{equation} where $g_{AB}$ is the metric on the $\xi_R=\mbox{ct.}$ hypersurfaces, and $dS_3$ is the line element on a unit ``radius'' (2+1)-dimensional de Sitter space, \begin{equation} d\Omega_{dS_3} = -d\chi_R^2 + \cosh^2\chi_R \,d\Omega_{S_2}. \end{equation} The Milne and the Rindler coordinates are related by analytic continuation, \begin{equation} \chi_R = r-i\pi/2 \hskip 1cm \xi_R = i t. \end{equation} Notice that $t$ is timelike inside the lightcone and becomes spacelike after analytical continuation to the outside, whereas $r$ is spacelike inside the lightcone but its analytical continuation is time-like. \begin{figure}[t] \centering \leavevmode\epsfysize=9cm \epsfbox{minkowski.eps}\\[3mm] \begin{quote} {\sl \caption[fig2]{\label{minkowski} Conformal diagram of Minkowski spacetime. The Milne coordinates $(t,r)$ cover the region inside the lightcone emanating from the origin $O$. The Rindler coordinates $(\xi_R,\chi_R)$ cover the outside of this lightcone. The thicker solid line in the central diamond shaped region corresponds to the position of the bubble wall.}} \end{quote} \end{figure} \section{Quantum state} Here we will consider two simple models. First we shall consider a massless field living in the Vilenkin-Ipser-Sikivie spacetime\cite{VIS}. The VIS spacetime represents the global gravitational field of a reflection symmetric domain wall, and can be constructed by gluing two Minkowski spaces at some $\xi_R=R_0$, the locus corresponding to the evolution of the bubble wall (see Fig.~\ref{VIS}). The second model we will study is a field which interacts with the tunneling field only on the bubble wall. For the tunneling field, we will assume the thin bubble wall approximation. More general models of the form (\ref{wish}) will be considered elsewhere\cite{next}. The quantization will be performed in the ``Rindler wedges'' of these spaces, because the hypersurfaces $\chi_R=\text{ct.}$ are Cauchy surfaces for the whole spacetime. \subsection{VIS model} \begin{figure}[t] \centering \leavevmode\epsfysize=8cm \epsfbox{VIS.eps}\\[3mm] \begin{quote} {\sl \caption[fig3]{\label{VIS} Conformal diagram of the VIS spacetime. This spacetime, which corresponds to the global gravitational field of a reflection symmetric domain wall, is constructed by identifying two Minkowski spacetimes at some $\xi_R=R_0$.}} \end{quote} \end{figure} Here we consider a massless field living in a spacetime constructed by gluing two Minkowski spaces at some $\xi_R=R_0$. We take Rindler coordinates in the region outside the origin of the two pieces, using a Rindler patch for each one. On each side, the Rindler coordinate $^{(l/r)}\xi_R$ , where the index $l$ or $r$ refers to the left or right pieces, ranges from $^{(l/r)}\xi_R=0$ on the lightcone to some value $^{(l/r)}\xi_R=R_0$, where the two Minkowski pieces are identified. Defining $^{(l)}\xi_R = R_0e^{\eta}$ and $^{(r)}\xi_R = R_0 e^{-\eta}$, we can coordinatize both pieces letting $\eta $ range from $-\infty$ to $\infty$. Then the line element outside the lightcone becomes \begin{equation} ds^2 = a(\eta)^2\left(d\eta^2-d\chi_R^2+\cosh^2\chi_Rd\,\Omega_2\right), \end{equation} where $a(\eta)=R_0e^{\eta}\theta(-\eta) + R_0e^{-\eta}\theta(\eta)$. Here $\theta(x)$ is the Heaviside step function. In order to construct a quantum state, we expand the field operator $\widehat{\bf\Phi}$ in terms of a sum over a complete set of mode functions times the corresponding creation and annihilation operators, \begin{equation} \widehat{\bf\Phi}=\sum_{plm}a_{plm}\Phi_{plm} + \mbox{h.c.}, \end{equation} The mode functions $\Phi_{plm}$ satisfy the field equation \begin{equation} \Box\Phi_{plm}=0\label{eqmassless}, \end{equation} where $\Box$ stands here for the four dimensional d'Alambertian operator in the VIS spacetime. Taking the ansatz \begin{equation} \Phi_{plm} = \frac{F_p(\eta)}{a(\eta)}{\mathcal Y}_{plm}(\tilde x) \end{equation} where $\tilde x = (\chi_R,\Omega) $, $\Omega=(\theta,\varphi)$, equation (\ref{eqmassless}) decouples into \begin{align} ^{dS}\Box {\mathcal Y}_{plm} &= (p^2+1){\mathcal Y}_{plm}\label{eqY}\\ \left[-\frac{d^2}{d\eta^2}-2\delta(\eta)\right]F_p &= p^2 F_p.\label{eqF} \end{align} Here $^{dS}\Box$ stands for the covariant d'Alembertian on a (2+1) de Sitter space. Equations (\ref{eqY})-(\ref{eqF}) have the interpretation that ${\mathcal Y}_{plm}$ are massive fields living in a (2+1) de Sitter space, with the mass spectrum given by the eigenvalues of the Schr{\"o}dinger equation for $F_p$. Solving (\ref{eqF}), we find that the spectrum has a continuous two-fold degenerate part for $p^2>0$ and a bound state with $p^2=-1$ (a zero mode). If we let $p$ take positive and negative values, the normalized mode functions $F_p$ for $p^2>0$, which are the usual scattering waves, can be written as \begin{equation} F_p = \frac{1}{\sqrt{2\pi}}\left((e^{ip\eta}+\rho(p) e^{-ip\eta})\theta(-\mbox{sgn}(p)\,\eta)\ + \sigma(p)e^{ip\eta}\theta\left(\mbox{sgn}(p)\,\eta\right)\right),\label{scatt} \end{equation} where \begin{align} \rho(p) &= -\frac{1}{i|p|+1},\\ \sigma(p) &= \frac{i|p|}{i|p|+1}. \end{align} The normalized supercurvature mode $p^2=-1$ is given by \begin{equation} F_{,-1} = \frac{a(\eta)}{R_0}, \end{equation} where the coma indicates that $-1$ refers to $p^2$ instead of $p$. As we are interested in a SO(3,1) invariant state, the natural choice for ${\mathcal Y}_{plm}$ are the positive frequency (2+1) Bunch-Davies modes \cite{bunchdavies}, \begin{equation} {\mathcal Y}_{plm}(\tilde x) = \sqrt{\frac{\Gamma(l+1+ip)\Gamma(l+1-ip)}{2}} \frac{P_{ip-1/2}^{-l-1/2}(i \sinh \chi_R)} {\sqrt{i \cosh \chi_R}}Y_{lm}(\Omega), \end{equation} where $Y_{lm}(\Omega)$ are the usual spherical harmonics. With this choice, it is straightforward to show that the quantum state for $\widehat\Phi$ is SO(3,1) invariant. Now we proceed to compute the two-point Wightman function $G^{(+)}(x,x')$, \begin{align} G^{(+)}(x,x') &\equiv \langle0|\widehat{\bf\Phi}(x)\widehat{\bf\Phi}(x') |0\rangle =\sum_{lm}\Phi_{-1,lm}(x)\overline{\Phi_{-1,lm}(x')} \nonumber\\&+ \int_{-\infty}^{\infty}dp\sum_{lm}\Phi_{plm}(x) \overline{\Phi_{plm}(x')}. \end{align} From now on we will suppose that the two points $x$ and $x'$ belong to the Rindler wedge of the ``left Minkowski'' space, so we will omit the $(l)$ index for notational simplicity. Direct substitution of the mode functions gives \begin{align} G^{(+)}(x,x') &= \frac{1}{R_0^2}\sum_{lm}{\mathcal Y}_{-1,lm}(\tilde x)\overline{{\mathcal Y}_{-1,lm}(\tilde x')}\nonumber\\&+ \frac{1}{2\pi R_0^2}\int^{\infty}_{-\infty}dp \left(\xi_M^{ip-1}\xi'_{M}{}^{-ip-1}\nonumber\right.\\&+\left. \frac{1}{ip-1}\xi_M^{ip-1}\xi'_{M}{}^{ip-1}\right)\sum_{lm} {\mathcal Y}_{plm}(\tilde x)\overline{{\mathcal Y}_{plm}(\tilde x')}, \end{align} where we have defined $\xi_M=e^\eta=\xi_R/R_0$. The two-point function $G^{(+)}$ is SO(3,1) invariant because is a sum of SO(3,1) invariant terms. Due to our choice of positive frequency modes, the $lm$ sums correspond to the two-point Wightman functions in the Euclidean vacuum for massive and a massless scalar fields living in $(2+1)$ de Sitter spacetime. The (3+1)-dimensional Lorentz group SO(3,1) is the same as the group of (2+1)-dimensional de Sitter transformations, so the two-point functions are Lorentz invariant by construction\footnote{\label{zeroinv}We will follow\cite{zeromodeA,zeromodeB,zeromodeC,jaumekirsten} to construct an SO(3,1) invariant state for the supercurvature massless mode with $p^2=-1$.} (its explicit form is given below). $G^{(+)}$ also depens on the quantity $\xi_R$. This is a function of the interval in Minkowski space time, so it is Lorentz invariant too. \begin{figure}[t] \centering \leavevmode\epsfysize=6cm \epsfbox{deSitter.eps}\\[3mm] \begin{quote} {\sl \caption[fig6]{\label{deSitter} Conformal diagram of a (2+1) de Sitter hyperboloid $\xi_R$=ct. Without loss of generality, we can take the point $\tilde x$ to lie in the ``origin'' O. Then $Z>1$ if $\tilde x'$ is timelike related with the origin, and $-1<Z<1$ if $\tilde x'$ is spacelike related with the origin and can be joined with it by means of a geodesic. If $Z<-1$, $\tilde x'$ is spacelike related with the origin but there are no geodesics connecting it with the origin. The function $\epsilon(x,x')$ is introduced in order to take into account the time ordering of those points which are timelike related. If this is the case, it evaluates to $\varepsilon$ if $\chi_R>\chi_R'$ and to $-\varepsilon$ if $\chi_R<\chi_R'$, where $\varepsilon$ is a small positive value. Two possible paths for $\chi'_R$ which pass through the origin are drawn (see discussion in Fig.~\ref{Zplane})}} \end{quote} \end{figure} First we will compute the contribution of the continuum. The $lm$ sum has been explicitly computed \cite{sumaY}, \begin{align} G^{(+)}_p(\tilde x,\tilde x') &\equiv \sum_{lm}{\mathcal Y}_{plm}\overline{{\mathcal Y}_{plm}} \nonumber\\&= \frac{\Gamma(1+ip)\Gamma(1-ip)}{(4\pi)^{3/2} \Gamma(3/2)} {}_2F_1\left[1+ip,1-ip;\frac{3}{2};{\frac{1+Z-i\epsilon}{2}}\right] \label{sumcont}, \end{align} where $Z(\tilde x,\tilde x')\equiv X^\mu(\tilde x)X_\mu(\tilde x') = -\sinh\chi_R\sinh\chi_R'+ \cosh\chi_R\cosh\chi_R'\cos\widehat{\Omega\Omega'}$, which is explicitly Lorentz invariant. Here $X^\mu(\tilde x)$ is the position of the point $\tilde x$ in the (3+1) Minkowski space where the $(2+1)$ de Sitter space is embedded as a timelike hyperboloid. The function $\epsilon(\tilde x,\tilde x')$ has been introduced to indicate at which side of the cut the hypergeometric function should be computed\footnote{The hypergeometric function in (\ref{sumcont}) has a branch cut along the real axis in the complex $Z$ plane from $Z=1$ to $Z=\infty$.}. It evaluates to $\varepsilon$ if $\tilde x$ and $\tilde x'$ are timelike related and $\chi_R>\chi_R'$, to $-\varepsilon$ if $\tilde x$ and $\tilde x'$ are timelike related and $\chi_R<\chi_R'$, and vanishes if $\tilde x$ and $\tilde x'$ are spacelike related, where $\varepsilon$ is a small positive constant (see Fig.~\ref{deSitter}). At the end of the calculation, we will take the limit $\varepsilon\to 0$. Introducing $\cos\tilde\zeta\equiv-\tilde Z \equiv -Z+i\epsilon$, the two-point function $G^{(+)}_p(\tilde x,\tilde x')$ can be compactly written as \begin{equation} G^{(+)}_p(\tilde x,\tilde x') = \frac{1}{4\pi\sin\tilde\zeta} \frac{\sinh p\tilde\zeta}{\sinh\pi p}. \end{equation} After performing the $p$ integration we obtain \begin{align} G^{(+)}_{\rm cont}(x,x') &= \frac{1}{8\pi^2\sigma}+\frac{1}{8\pi R_0^2}\left( 2\cot\tilde\zeta\left(\tilde\zeta - \mbox{arc tan}\frac{\sin\tilde\zeta}{\cos\tilde\zeta+\xi_M\xi_M'}\right)\right. \nonumber\\ &+\left. \log\frac{(\xi_M\xi'_M)^2} {\sin\tilde\zeta^2+(\cos\tilde\zeta+\xi_M\xi'_M)^2}\right). \end{align} \begin{figure}[t] \centering \leavevmode\epsfysize=4cm \epsfbox{Zplane.eps}\\[3mm] \begin{quote} {\sl \caption[fig6]{\label{Zplane} Paths in the complex $\tilde Z$- and $\tilde \zeta$-planes for the curves shown in Fig.~\ref{deSitter}, where we hold the point $\tilde x$ fixed at O while moving $\tilde x'$ around the $(2+1)$ de Sitter space. If $Z<-1$, then $\tilde\zeta$ is purely imaginary. If $-1<Z<1$, $\tilde \zeta$ is essentially real (if it were not for the small $i\epsilon$ imaginary part). In this case, if $\chi_R<\chi_R'$, $-\pi<\tilde \zeta<0$, but if $\chi_R>\chi_R'$ then $0<\tilde\zeta<\pi$. The coincidence limit corresponds to both $\tilde\zeta=\pm \pi$, depending on that we approach $\tilde x$ from ``abov'' or ``below''. When $\tilde x$ and $\tilde x'$ are timelike related, $\tilde\zeta$ has both imaginary and real parts. Its real part is $\pm\pi$ depending on whether $\chi_R$ is greater or less than $\chi_R'$, respectively.}} \end{quote} \end{figure} The ``supercurvature'' contribution of the $p^2=-1$ mode is in fact divergent. This is related to the zero mode problem of massless quantum fields in spacetimes with compact Cauchy surfaces. Following the usual prescription \cite{zeromodeA,zeromodeB,zeromodeC,jaumekirsten}, we formally write this divergent term as a divergent piece plus a finite one, \begin{equation} G^{(+)}_{\rm sup}(x,x') = \frac{1}{4\pi^2}\langle0|Q^2|0\rangle + \sum_{l>0,m}\frac{1}{R_0^2}{\mathcal Y}_{-1,lm}\overline{{\mathcal Y}_{-1,lm}} , \end{equation} where the infinity has been hidden in an infinite constant (see \cite{jaumekirsten} for details). After, when taking derivatives to compute the energy-momentum tensor, this divergent constant term will give no contribution. The sum can be performed, and the result is \begin{align} G^{(+)}_{\rm sup}(x,x') &=\frac{1}{4\pi^2}\langle0|Q^2|0\rangle + \frac{1}{8\pi^2R_0^2}\left( - 2\tilde\zeta\cot\tilde\zeta\right.\nonumber\\&+\left. (2\chi+i\pi)\tanh\chi_R+(2\chi'-i\pi)\tanh\chi'_R\frac{}{}\right), \end{align} where we have dropped an irrelevant constant. Adding the continuum and supercurvature contributions, and symmetrizing the result with respect to $x$ and $x'$, we finally find the symmetric Hadamard two-point function (for pairs of points $x$, $x'$ outside the lightcone in the ``left'' Minkowski patch), \begin{align} G^{(1)}(x,x') &=\frac{1}{4\pi^2\sigma}+\frac{1}{4\pi^2R_0^2} \left(2\chi_R\tanh\chi_R+ 2\chi'_R\tanh\chi'_R\nonumber\right.\\ &- \left.2\,\cot\tilde\zeta\,\mbox{arc tan}\frac{\sin\tilde\zeta}{\cos\tilde\zeta+\xi_M\xi_M'}\nonumber\right.\\ &+\left. \log\frac{(\xi_M\xi'_M)^2} {\sin^2\tilde\zeta+(\cos\tilde\zeta+\xi_M\xi'_M)^2}\right)+ \frac{1}{2\pi^2}\langle0|Q^2|0\rangle \nonumber\\ &= \frac{1}{4\pi^2\sigma}+W(x,x'), \end{align} where \begin{equation} \sigma=\frac{1}{2}\left(\xi_R^2+\xi'_{R}{}^2+ 2\xi_R\xi'_R\cos\tilde\zeta\right) \end{equation} is one half of the square geodesic distance in flat spacetime. The first term in the final expression for $G^{(1)}(x,x')$ is the usual Minkowski ultraviolet divergence. The second term, $W(x,x')$, is due to the nontrivial geometric boundary conditions imposed by the symmetry of our problem. If $W(x,x')$ were not singular, the state would be of the Hadamard form\cite{hadamardA,hadamardB,hadamardC}. But $W(x,x')$ has local and nonlocal singularities. In the coincidence limit, it is divergent on the bubble wall. It is logarithmically singular whenever one of the points is on the lightcone emanating from the origin. It is also singular when $x$ and $x'$ satisfy the relation $\sin^2\tilde\zeta+(\cos\tilde\zeta+\xi_M\xi'_M)^2=0$, so the argument of the logarithm diverges. The roots of this equation are at \begin{equation} \xi_M^{\rm s}{\xi'}^{\rm s}_M=-e^{\pm i\tilde\zeta_{\rm s}}.\label{rootsing} \end{equation} To clarify the position of the singularities, let us fix the point $x_{\rm s}$ and look for the points $x'_{\rm s}$ which make $G^{(+)}(x_{\rm s},x'_{\rm s})$ singular. Taking into account that $\xi_M^{\rm s}$ and ${\xi'}^{\rm s}_M$ should be real and should satisfy $0<\xi_M^{\rm s},{\xi'}^{\rm s}_M<1$, it is seen from (\ref{rootsing}) that the allowed values of $\tilde\zeta_{\rm s}$ are of the form $\pm\pi+iy$ (i.e., $\tilde x_{\rm s}$ and $\tilde x'_{\rm s}$ are ``timelike'' separated on a (2+1) de Sitter hyperboloid, see Fig.~\ref{Zplane}), with $y>-\log\xi_M^{\rm s}>0$. Without loss of generality, we can assume that $\tilde x_{\rm s}= (0,0,0)$. Then $\cos\tilde\zeta_{\rm s}=-\cosh y=-\cosh {\chi_R'}^{\rm s}\cos\theta'_{\rm s}$. This implies that $0\leq\theta'_{\rm s}\leq\pi/2$, and we have no restriction on $\varphi'_{\rm s}$. Since $\cosh {\chi_R'}^{\rm s} = \cosh y/\cos\theta'_{\rm s}\geq \cosh y$, we find that ${\chi_R'}^{\rm s}\geq y\geq-\log\xi_M^{\rm s}$ or ${\chi_R'}^{\rm s}\leq-y\leq\log\xi_M^{\rm s}$. So the region inside of which (for any value of $\Omega'$) $G^{(+)}(x,x')$ is non singular (apart from the singular points on the lightcone from $x$) is limited by the curves \begin{figure}[t] \centering \leavevmode\epsfysize=6cm \epsfbox{singularities.eps}\\[3mm] \begin{quote} {\sl \caption[fig6]{\label{singularities} Nonlocal singularities in the upper half ``left Minkowski'', in the VIS model. Inside the shadowed region the two-point function $G^{(1)}(x,x')$, considered as a function of $x'$ with $x$ fixed, is singular only on the lightcone from $x$.}} \end{quote} \end{figure} \begin{align} {\xi'}^{\rm n.s.}_M&=\frac{1}{e^y\xi_M^{\rm s.}}\label{singularA}\\ {\chi'}^{\rm n.s.}_R &= \pm y\label{singularB}, \end{align} with $y>-\log\xi_M^{\rm s}$, the lightcone from the origin and the bubble wall (the superscript n.s. stands for ``nearest (nonlocal) singularity'', see Fig.~\ref{singularities}). Note that as $x$ approaches to the bubble wall (i.e., $\xi_M\to 1$), the distance to the nearest singular point $x'$ reduces. Consistently, in the limiting case when $x$ is on the wall, $W(x,x')$ is singular on the coincidence limit. Let us now check the causal relationship between singular points satisfying equation (\ref{rootsing}). If we compute $\sigma(x_{\rm s},x'_{\rm s})$, we will find \begin{equation} \sigma(x_{\rm s},x'_{\rm s})=\frac{e^{-2 y}}{2\xi_M^{{\rm s}\,2}}(1-\xi_M^{{\rm s}\,2}) (1-\xi_M^{{\rm s}\,2} e^{2y})\leq 0, \end{equation} where the last inequality follows from $y\geq-\log\xi_M^{\rm s}$, $0\leq\xi_M^{\rm s}\leq 1$. The equality can only be realized if $x_{\rm s}$ is on the bubble wall. Then, in this case, there exist nonlocal singularities (of $W(x,x')$) which are null related. But if $x_{\rm s}$ is not on the bubble wall, its singular partners are always time-like related with it. Summarizing, the two-point function is locally Hadamard everywhere except on the bubble wall and on the lightcone. Moreover, it has (harmless, see discussion below) nonlocal singularities. \subsection{ST model} In this second model, which has previously been considered by Sasaki and Tanaka \cite{TMdecay}, the $\widehat\Phi$ field interacts with the tunneling field $\sigma$ only on the bubble wall. We assume the infinitely thin-wall approximation, so the interaction term can be written as \begin{equation} m^2(\eta) = 2 \frac{V_0}{R_0^2} \delta(\eta), \end{equation} where $V_0>0$ characterizes the strength of the interaction and $R_0$ is the radius of the bubble wall. Decomposing the field $\widehat{\bf \Phi}$ as before, we find that the Schr{\"o}dinger equation for $F_p$ takes the form \begin{equation} -F''_p+2 V_0\delta(\eta)F_p = p^2 F_p, \end{equation} Now the spectrum is purely continuous with $p^2>0$. The solution of this Schr{\"o}dinger equation is the scattering basis (\ref{scatt}) with the transmission and reflection coefficients given by \begin{align} \rho(p) &= \frac{V_0}{i|p|-V_0},\\ \sigma(p) &= \frac{i|p|}{i|p|-V_0}. \end{align} Following a similar path\footnote{Detailed computations will be presented elsewhere\cite{next}}, we arrive at the following Hadamard two-point function (for points $x$, $x'$ in the Rindler wedge and inside the bubble), \begin{align} G^{(1)}(x,x')&=\frac{1}{4\pi^2\sigma}+ \frac{1}{4\pi^2 i}\frac{1}{\xi_R\xi'_R\sin\tilde\zeta} \left(_2F_1\left[1,V_0;V_0+1;-e^{i\tilde\zeta} \frac{\xi_R\xi'_R}{R_0^2}\right]\right.\nonumber\\ &- \left._2F_1\left[1,V_0;V_0+1;-e^{-i\tilde\zeta} \frac{\xi_R\xi'_R}{R_0^2}\right]\right)=\frac{1}{4\pi^2\sigma}+W(x,x'), \end{align} which is explicitly SO(3,1) invariant. If we take the coincidence limit, the function $W(x,x')$ has divergences on the bubble wall, so the state is not locally Hadamard. Apart from this, it has nonlocal logarithmic singularities at the points where the argument of the hypergeometric functions become 1, i.e., whenever $\xi_M\xi_M'=-\exp(\pm i\tilde \zeta)$. This is the same relation we found in the VIS model. Borrowing the conclusions from the VIS model, the state is locally Hadamard everywhere except on the bubble wall, and has (harmless) timelike nonlocal singularities (except also on the bubble wall). As we have seen, the two models we have considered share two singular behaviors: the existence of nonlocal singularities and the singularity of $W(x,x')$ in the coincidence limit on the bubble wall. These singularities seem to be related with the oversimplification of the model. Presumably, if instead of a $\delta$-like term interaction we had introduced a smooth function, these divergences would disappear. \section{Renormalized expectation value of the stress tensor} As we have pointed out, for the two models we have studied the singularities of $G^{(1)}(x,x')$ are nearly of the Hadamard type. We can use the point-splitting regularization prescription to compute the renormalized expectation value of the stress-energy momentum tensor \cite{christensen,waldreg,folacci}, \begin{align} \langle T_{ab}\rangle &= \frac{1}{2}\lim_{x\to x'}{\mathcal D}_{ab'}[W(x,x')],\\ {\mathcal D}_{ab'} &= \nabla_a\nabla_{b'}- \frac{1}{2}g_{ab'}g^{cd'}\nabla_c\nabla_{d'}. \end{align} Noticing that $\cos\tilde\zeta = -Z+i\epsilon=-\cos\sqrt{2\,\,^{dS}\sigma}+i\epsilon$, where $^{dS}\sigma$ is one half of the square distance in a unit (2+1) de Sitter spacetime, the covariant derivatives in the ``de Sitter'' direction are easily computed from \cite{christensen,hadamardC} \begin{align} [^{dS}\sigma_{;A}] &=0, \\ {[^{dS}\sigma_{;AB'}]} &= -\frac{g_{AB}}{\xi_R^2}, \end{align} where the brackets stand for the coincidence limit. \subsection{VIS model} The renormalized expectation value of the stress tensor turns out to be \begin{align} \langle T_{\xi_R\xi_R}\rangle&=\frac{\xi_R^2-2 R_0^2}{4\pi^2 R_0^2 (R_0^2-\xi_R^2)^2}, \\ \langle T_{AB}\rangle &= -\frac{\xi_R^4-3 R_0^2\xi_R^2+6 R_0^4} {12\pi^2R_0^2(R_0^2-\xi_R^2)^3}\,g_{AB}. \end{align} where $0\leq\xi_R\leq R_0$ (i.e., $x$ is in the left Rindler wedge and inside the bubble). It is clear from the expression that the energy-momentum tensor behaves somewhat better than the two-point function. It is divergent on the bubble wall, but behaves smoothly on the lightcone. So it can be analytically continued to the inside of the lightcone. There it behaves like the energy-momentum tensor of a perfect fluid, \begin{align} \langle T_{ab}\rangle = (\rho + p)u_a u_b + p g_{ab}, \end{align} where $u^a=(\partial_t)^a$. On the lightcone it satisfies the equation of state $p=-\rho$ whith \begin{align} \rho=\frac{1}{2\pi R_0^4} \end{align} For large $t$ the equation of state turns out to be \begin{equation} p=-\frac{1}{3}\rho, \end{equation} with \begin{equation} \rho=\frac{1}{4\pi^2}\frac{1}{R_0^2t^2}. \end{equation} Taking into account that the field is massless except on the bubble-wall, one might naively expect that the energy momentum tensor would behave like radiation, with $\rho\propto t^{-4}$. Instead of this, we have found that it decreases slower. In fact, it can be shown that its behaviour is dominated by gradients of the supercurvature modes\footnote{The particle content and interpretation of the vacuum we have considered will be discussed elsewhere\cite{next}\label{particle}}. \subsection{ST model} For the ST model, we find \begin{align} \langle T_{\xi_R\xi_R}\rangle &= \frac{1}{2\pi^2R_0^4} \frac{V_0}{V_0+2}{}_2F_1\left[3,V_0+2;V_0+3; \left(\frac{\xi_R}{R_0}\right)^2\right],\\ \langle T_{AB}\rangle &= \frac{1}{2\pi^2R_0^4}\frac{V_0}{V_0+2}\left( _2F_1\left[3,V_0+2;V_0+3; \left(\frac{\xi_R}{R_0}\right)^2\right]\nonumber\right.\\ &+\left.2\left(\frac{\xi_R}{R_0}\right)^2 \frac{V_0+2}{V_0+3}{}_2F_1\left[4,V_0+3;V_0+4; \left(\frac{\xi_R}{R_0}\right)^2\right]\right)g_{AB}, \end{align} where $0\leq\xi_R\leq R_0$. As before, the energy-momentum tensor turns out to be singular only on the bubble wall\footnote{The quantum state found in\cite{TMdecay} has the problem of being ill defined on the light-cone. This singularity propagates to the renormalized energy-momentum tensor, making it to blow up on the light cone. This seems to be due to an inappropiate normalization of the mode functions.}. Continuing analytically the results to the inside of the light-cone, we find that it is of the perfect fluid form. On the lightcone it satisfies the equation of state $p=-\rho$ with \begin{align} \rho=-\frac{1}{2 \pi^2 R_0^4}\frac{V_0}{V_0+2}. \end{align} For large $t$ the equation of state turns out to be \begin{equation} p=\rho, \end{equation} with \begin{equation} \rho = -\frac{4\,R_0^2}{t^6}. \end{equation} Notice that in this model the energy density is negative and decreases faster than radiation\footnotemark[5]. \section{Discussion} In this paper we have performed the computation of $\langle T_{ab}\rangle$ in a quantum state which fulfills our basic requirement of SO(3,1) invariance. In fact, we have just outlined the most simple method to find a SO(3,1) invariant state. The question is whether by choosing a different set of modes we can also obtain an inequivalent SO(3,1) invariant state but also of the Hadamard form. A theorem due to Kay and Wald \cite{waldkay} is illuminating in this respect. The theorem states that in a spacetime with a bifurcate Killing horizon there can exist at most one regular quasifree state invariant under the isometry which generates the bifurcate Killing horizon. Let us briefly analyze the conditions under which the theorem holds. In (3+1) spacetimes, we get a bifurcate Killing horizon whenever a one parameter group of isometries leaves invariant a 2-dimensional spacelike manifold $\Sigma$. The bifurcate Killing horizon is generated by the null geodesics orthogonal to $\Sigma$\cite{waldkay}. For example, Minkowski spacetime has bifurcate Killing horizons. The isometry group is a one-parameter subgroup of Lorentz boosts, and the manifold $\Sigma$ is a two-plane. Any SO(3,1) invariant spacetime, where the line element can be written in the form \begin{equation} ds^2 = d\xi_R^2 + a(\xi_R)^2(-d\chi_R^2+\cosh^2\chi_Rd\Omega), \end{equation} has a SO(3,1) invariant bifurcate Killing horizon. Noticing that the $\xi_R=\mbox{ct.}$ hypersurfaces are (2+1) de Sitter spaces which can be thought as embedded in a (3+1) Minkowski space, any boost generator on these hypersurfaces is the infinitesimal generator of a isometry which 1) leaves invariant a spacelike 2-manifold (so we get a bifurcate Killing horizon) and 2) leaves any SO(3,1) symmetric state invariant. We can take, for example, the boost generator in the $ZT$ plane of the embedding Minkowski space. Expressed in the Rindler coordinates, it becomes \begin{equation} \xi^a = -\cos\theta\frac{\partial}{\partial\chi_R} + \tanh\chi_R\sin\theta \frac{\partial}{\partial\theta}. \end{equation} The Killing field $\xi^a$ leaves invariant the spacelike 2-manifold $\theta=\pi/2$, $\chi_R=0$. All bubble spacetimes with or without the inclusion of gravity do possess this bifurcate Killing horizon. A (pure) quasifree ground state is the (wider) algebraic version (see \cite{waldkay} and references therein) of what is usually called a ``frequency splitting'' Fock vacuum state. A quasifree state has the special property of being completely characterized from its two-point function. A regular\footnote{We include the notion of globally Hadamard in the definition of a regular state.} quasifree ground state is a quasifree ground state whose two-point symmetric function is globally Hadamard and which has no zero modes. The VIS model has a zero mode (as any massless field in spacetimes with compact Cauchy surfaces have \cite{waldkay}), so the theorem cannot be directly applied. Also strictly speaking, the quantum state we have found for the ST model does not fulfill the requirements of the theorem, because it is not globally Hadamard. Roughly speaking, a two-point function is said to be globally Hadamard if it is locally Hadamard and in addition has nonlocal singularities only at points $x$, $x'$ which are null related within a causal normal neighborhood of a Cauchy hypersurface\footnote{A more rigorous definition of globally Hadamard states can be found in \cite{waldkay}}. As we have seen, if we ignore the problems on the bubble wall, the Hadamard function $G^{(1)}(x,x'$) we have found for the ST model has nonlocal singularities, but they are timelike related. So, if it were not for the singularities on the bubble wall, the state would be globally Hadamard and without zero modes. As stated before, we think that the singularities on the bubble wall would disappear if the potential were modeled by a smooth function instead of by a $\delta$-like term, making the state globally Hadamard. Then, symmetry would suffice to determine the (physically admissible) quantum state for this model. Generic models which would not present these pathologies will be presented elsewhere\cite{next} \section*{Acknowledgements} I would like to thank Edgard Gunzig for his kind hospitality at the Peyresq-3 meeting. I am grateful to Jaume Garriga for many helpful discussions. I acknowledge support from European Project CI1-CT94-0004, and from CICYT under contract AEN98-1093.
\section{Introduction} Low mass X--ray binaries, where a stellar mass black hole or neutron star accretes matter from its Roche Lobe filling companion, are generally {\it transient\/} systems. The mass accretion rate is low enough for the accretion disc to be unstable, leading to long quiescence periods followed by dramatic outbursts (e.g.\ King et al.\ 1997a; King, Kolb \& Szuszkiewicz 1997b). For the black hole systems, the outburst often shows a rapid rise from a very faint quiescent state, reaching luminosities close to the Eddington limit in a course of a few days. The outburst then declines roughly exponentially, with a characteristic time scale of 30--40 days in both the X--ray and optical flux (e.g.\ Chen, Schrader \& Livio 1997). During the decline the X--ray spectra and variability go through a sequence of well defined states. For luminosities close to the Eddington limit, $L_{\rm Edd}$, the spectrum has both a strong soft component from the accretion disc and a strong power law tail (the very high state). At lower luminosities the hard tail decreases and steepens, so the spectrum is dominated by the soft component (the high state). There is then a dramatic transition, where the soft component decreases substantially in temperature and luminosity, and the spectrum is instead dominated by a hard power law tail (the low state) (see e.g.\ Tanaka \& Lewin 1995; van der Klis 1995; Tanaka \& Shibazaki 1996). {GS~2023+338}\ broke with this pattern in several ways. Firstly, it never clearly showed the very high or high state spectrum even though its luminosity was close to the Eddington limit. Secondly, it showed dramatic flux and spectral variability, part of which can be attributed to a heavy and strongly variable photo--electric absorption (Tanaka \& Lewin 1995; Oosterbroek et al.\ 1997). Thirdly, its optical spectrum was unlike that from any other transient system, with strong, broad lines from H, HeI and HeII: normally these lines are rather weak (Casares et al.\ 1991). In this paper we analyze in detail archival {\it Ginga}\ data of {GS~2023+338}\ covering the beginning of its outburst, 23--30 May 1989, attempting to deconvolve the primary spectrum from effects of subsequent reprocessing (absorption, Compton reflection, line fluorescence). The spectrum at the peak luminosity is very different to that seen in more canonical transient systems such as Nova Muscae 1991. We postulate that this spectrum represents super Eddington accretion rates (Inoue 1993), and speculate that this and many of the other peculiar properties of {GS~2023+338}\ are due to the wide separation of the binary, as suggested for GRO J1655--40 (Hynes et al.\ 1998). The accumulating accretion disc is then very large, so that when the disc instability is triggered there is much more mass in the quiescent disc than in a more typical transient. The accretion can then be super Eddington, giving rise to some form of strong mass outflow which manifests itself as heavy photo--electric absorption, and produces the intense optical line emission. This absorption shrouds the source as the accretion rate declines below Eddington, masking its transition to the standard very high spectrum expected at these luminosities. As the outflow expands with time, the absorption becomes less extreme. Eventually, there are times when the intrinsic source spectrum is not obscured, but by this point the mass accretion rate has declined so that the source is in the low/hard state. \section{Data reduction and background subtraction} Data were extracted from the UK {\it Ginga}\ data archive at Leicester University and reduced in the usual manner. Background subtraction poses a problem for a source as bright and hard as {GS~2023+338}\ since the background monitors (most importantly the Surplus above Upper Discriminator, or SUD) are contaminated by counts from the source. We have used the same method as described in \.{Z}ycki, Done \& Smith (1999; hereafter Paper I) to recover the uncontaminated SUD values and then we used the 'universal' background subtraction method (Hayashida et al.\ 1989) to estimate and subtract the background. We assume 0.5 per cent systematic errors in the data unless stated otherwise. \begin{figure} \epsfxsize=0.5\textwidth \epsfbox[10 190 590 670]{lc_1-30k.ps} \caption{{\it Ginga\/} light--curve (count rate in 1--30 keV; 2 sec time bins) on 30th May, when the source's flux reached its maximum and apparently saturated as well as showed dramatic variability. The count rate was corrected for background, dead-time and aspect. Beginning of good data, $t_0$, is 4:31:13. Data for the unabsorbed spectrum (Section~\ref{sec:may30}; Figure~\ref{fig:may30spec}; Table~\ref{tab:may30spec}), and for the absorbed one shown in Figure~\ref{fig:s329_abs} were extracted from time interval $i-1$. Spectra from time intervals $i-4$ and $i-5$ are analyzed in Section~\ref{sec:soft}. They are consistent with soft/high state spectra of SXT, when corrected for absorption. \label{fig:may30_lc}} \end{figure} \begin{figure} \epsfxsize=0.5\textwidth \epsfbox[10 180 600 690]{colcol.ps} \caption{ Colour--colour plot for data obtained on 30th May. Time bin is 1 second. The circle marks the region where data for the unabsorbed spectrum were extracted from (Section~\ref{sec:may30}), while the rectangle marks similar region for the absorbed spectrum shown in Figure~\ref{fig:s329_abs}. Contours labeled $i-4$ and $i-5$ mark approximate regions where the soft state data were extracted from (cf.\ Figure~\ref{fig:may30_lc}; Section~\ref{sec:soft}). Very rapid and chaotic variability was observed during time interval $i-7$, most likely due to photo-electric absorption (Section~\ref{sec:heavyabs}). For comparison, the two big dots show position of Nova Muscae 1991 on January 11 and January 16 (see Section~\ref{sec:compar}, Figure~\ref{fig:gs_nm}). The crosses show typical error bars on the colours: the bigger one is for $i-7$, the smaller one for earlier observations. \label{fig:colcol}} \end{figure} \section{Models} We use a variety of models for spectral description. For the soft component we generally use the multi-temperature blackbody model (Mitsuda et al.\ 1984). For simple estimates we use the simple version implemented as {\tt diskbb} in {\sc XSPEC}. This however does not include the colour temperature correction (e.g.\ Shimura \& Takahara 1995), or the torque-free boundary condition term in the expression for temperature, $1-\sqrt{6R_{\rm g}/r}$ (assuming Newtonian dynamics; Shakura \& Sunyaev 1973), where $R_{\rm g}\equiv G M/c^2$. We implemented these corrections in a model called {\tt diskspec} (see also Gierli\'{n}ski et al.\ 1998). The main parameter can be chosen to be either the temperature at the inner disc radius or the ratio of mass accretion rate and the mass of the central object. \begin{table*} \caption{Model fitting of the May 30.\ unabsorbed spectrum. \label{tab:may30spec}} \begin{tabular}{lccccc} parameter & units & 0 & A & B & C \\ \hline $N_{\rm H}$ & $10^{22}\,{\rm cm^{-2}}$ & 3.7 & $0.50^{+0.24}$ & $0.50^{+0.13}$ & $0.50^{+0.22}$ \\ $k T_{\rm soft}$ & keV & 0.23 & $0.270\pm 0.015$ & $0.322\pm 0.015$ & $0.281\pm 0.013$ \\ $D_{\rm in}$ & $R_{\rm g}$ & 370 & $71^{+15}_{-12}$ & $41 \pm 8$ & $62^{+13}_{-10}$ \\ $\Gamma$ & & 1.57 & $1.03\pm 0.04$ & $1.690^{+0.023}_{-0.011}\,^{a\,b}$& $1.700^{+0.013}_{-0.011}\,^{a\,b}$ \\ $E_c$ & keV & 60 & $19.6^{+1.2}_{-1.0}$& -- & -- \\ $k T_{\rm e}$ & keV & -- & -- & $9.2^{+0.4}_{-0.2}\,^b$ & $9.6^{+0.7}_{-0.5}\,^b$ \\ $\tau_{\rm es}$ & & -- & -- & -- & $6.49\pm 0.07\,^b$ \\ $T_0$ & keV & -- & -- & $0.83\pm 0.03$ & $1.44\pm 0.09$ \\ \hline $\Omega_{\rm r}$ & & -- & -- & $0.13^{+0.05}_{-0.03}$ & $0.17^{+0.16}_{-0.05}$ \\ $\xi$ & erg/cm s &-- & -- & $(2.5^{+7.5}_{-2.0})\times 10^3$ & $(6^{+18}_{-4})\times 10^3 $ \\ $R_{\rm in}$ & $R_{\rm g}$ & -- & -- & $6^{+1.1}$ & $6^{+2}$ \\ $E_{\rm edge}$& keV& -- & $7.70\pm 0.13$ & -- & -- \\ $\tau_{\rm edge}$& & -- & $1.19\pm 0.13$ & -- & -- \\ EW & eV & -- & $80^{+13}_{-20}$ & $67 \pm 16$ & $70 \pm 17 $ \\ $\chi^2/$dof & & 555/32 & 23.5/29 & 37.3/27 & 25.7/27 \\ \hline \end{tabular} \medskip Model 0: absorption*(disc blackbody + cutoff power law) Model A: absorption*smeared edge*( disc blackbody + cutoff power law + gaussian) Model B: absorption*(disc blackbody + comptonized blackbody ({\tt thComp}) + {\tt relrepr}\ + gaussian) Model C: absorption*(disc blackbody + comptonized disc blackbody ({\tt thComp}) + {\tt relrepr}\ +gaussian) $N_{\rm H}$ -- hydrogen column density. Its hard lower limit (interstellar value) is assumed 0.5.\\ $D_{\rm in}$ -- inner disc radius computed from the amplitude of the {\tt diskbb} model \\ $k T_{\rm e}$, $\tau_{\rm es}$ -- electron temperature and optical depth of the comptonizing cloud \\ $T_0$ -- temperature of the seed photon input spectrum for comptonization \\ $E_c$ -- e-folding energy in the cutoff power law model \\ $\Omega_{\rm r},\ \xi$ -- amplitude and ionization parameter of the reprocessed component \\ $R_{\rm in}$ -- inner disc radius determining the level of relativistic smearing \\ EW -- equivalent width of the additional gaussian line \\ $^{a}\,$ Asymptotic value of $\Gamma$ in the ST80 solution. \\ $^b\,$ $\Gamma$, $k T_{\rm e}$ and $\tau_{\rm es}$ are of course not independent. \\ Parameters' uncertainties are 90 per cent confidence limits for one interesting parameter i.e.\ $\Delta\chi^2=2.71$. \end{table*} We use an analytic thermal comptonization model to describe the hard component: {\tt thComp}, based on solution of the Kompaneets equation (Lightman \& Zdziarski 1987). For accurate modelling of inverse Compton spectra we also use a Monte Carlo simulation code. It is based on standard methods of simulations of the inverse Compton process as described in detail by Pozdnyakov, Sobol \& Sunyaev (1983) and G\'{o}recki \& Wilczewski (1984) (see Appendix A). The X--ray reprocessed component is modelled using the angle-dependent Green's functions of Magdziarz \& Zdziarski (1995) convolved with a given continuum model to produce the reflected continuum, with photo--electric opacities calculated by a simple photo-ionization code described in Done et al.\ (1992). For the iron fluorescent/recombination K$\alpha$ line we used the Monte Carlo code of \.{Z}ycki \& Czerny (1994). We constructed Green's functions for the line emission (i.e.\ emission for a monochromatic irradiation flux), as functions of ionization and iron abundance, which can be folded with a given continuum model, to compute total line profile, which is then added to the reflected continuum. Parameters of the model are: the amplitude of the total reprocessed component, defined as the solid angle of the reprocessor as seen from the X--ray source, $\Omega$, normalized to $2 \pi$, $\Omega_{\rm r}\equiv \Omega/2 \pi$ and the ionization parameter, $\xi \equiv 4\pi F_{\rm X}/n$, where $F_{\rm X}$ is the illuminating flux in the 5 eV -- 20 keV band and $n$ is the electron number density. We assume cosmic abundances of Morrison \& McCammon (1983) with the possibility of a variable iron abundance in the reprocessor. For more model details see Paper I. The total reprocessed spectrum can then be convolved with the {\sc XSPEC} {\tt diskline} model (Fabian et al.\ 1989; modified to include light bending in Schwarzschild metric) to simulate the relativistic smearing. The main parameter here is the inner radius of the disc, $R_{\rm in}$, assuming a given form of the irradiation emissivity, for which we adopt $F_{\rm irr}(r) \propto r^{-3}$ (see Paper I). The outer disc radius is fixed at $10^4\,R_{\rm g}$. This is smaller than the expected outer radius of the disc in {GS~2023+338}\ ($\sim 10^5\,R_{\rm g}$; see Section~\ref{sec:discuss}), but with our assumed $F_{\rm irr}(r)$ the fractional contribution from the ring $10^4$--$10^5\,R_{\rm g}$ would be negligible, $\approx 5\times 10^{-4}$. The total model will be referred to as {\tt relrepr}. Strong photo--electric absorption is modelled by either the simple formula $F \propto \exp(-\tau_{\rm eff})$ where $\tau_{\rm eff} = \sqrt{\tau_{\rm abs} (\tau_{\rm abs} + \tau_{\rm es})}$ (Rybicki \& Lightman 1979; called {\tt thabs}) or by a proper Monte Carlo transmission code for spherical geometry. We fix the inclination of the source at $i=56^{\circ}$ and assume the mass of the black hole is $12\,{\rm M}_{\odot}$ (Shahbaz et al.\ 1994). We assume the interstellar $N_{\rm H}=5\times 10^{21}\,{\rm cm^{-2}}$, corresponding to $A_V=3$ (Chen et al.\ 1997; Shahbaz et al.\ 1994). The spectral analysis was performed using {\sc XSPEC} ver.\ 10 (Arnaud 1996) into which all the non-standard models mentioned above were implemented as local models. \section{The unabsorbed spectrum} \label{sec:may30} On the 30th of May {\it Ginga}\ observed {GS~2023+338}\ in a very bright state (see Fig.~\ref{fig:may30_lc}). The observed source flux reached $6.5\times 10^{-7}$ erg/s/cm$^2$ (1 -- 40 keV; corresponding to $L_{\rm X}\approx 10^{39} {\rm erg\ s}^{-1}$ at $d=3.5$ kpc) and it apparently saturated at this level. For about 200 sec after the beginning of observation the source spectrum was very stable (Fig.~\ref{fig:colcol}) and it did not show any strong photo--electric absorption, so we used the 200 sec average spectrum for detailed analysis. \subsection{Phenomenological description} Generally, the data can be described (Table~\ref{tab:may30spec}) as a sum of a soft component and a hard cutoff power law component ($\Gamma\approx 1.0$) with e-folding energy of $E_{\rm c}\approx 20\,$keV (i.e.\ the cutoff is obviously visible in the {\it Ginga}\ band). Superimposed on this are spectral features near 6--10 keV. These features can be phenomenologically modelled by a narrow gaussian at 6.4 keV (EW=77 eV) and the smeared edge (Ebisawa 1991) with $E\approx 7.7\,$keV and $\tau\approx 1.2$ (Model A in Table~\ref{tab:may30spec}. The presence of the spectral features is highly significant; the best fit model using only two continuum components gives $\chi^2=555/32\,$dof (Model 0 in Table~\ref{tab:may30spec}). The soft component can be described equally well by both simple blackbody and disc blackbody model. This is due to the fact that only the Wien cutoff is visible in {\it Ginga}\ band. Assuming the model which would give the lowest bolometric correction i.e.\ a blackbody spectrum and assuming further that at that time there was no photo-electric absorption in excess of the interstellar value, the derived temperature is $k T_{\rm soft}=0.21\,$keV and the normalization of the component corresponds to the emitting area of $\approx 3\times 10^{6}\,{\rm km^2}\approx (90\,R_{\rm g})^2$. Extrapolated to lower energies, this component then contributes significantly to the total energy budget, with a total bolometric flux of $10^{-6}$ ergs cm$^{-2}$ s$^{-1}$. Thus the lowest possible luminosity is around the Eddington limit, but it is likely to be rather higher since the soft spectrum must be broader than a single blackbody and the absorption may be rather higher than the (poorly known) interstellar value of $5\times 10^{21}$ cm$^{-2}$. \begin{figure} \epsfxsize = 0.45\textwidth \epsfbox[0 190 550 720]{unabs_sp.ps} \caption{Source spectrum on 30 May, 4:31:13 -- 4:34:33 (200 sec), when the flux apparently saturated at very high level (Figure~\ref{fig:may30_lc}, part of time interval $i-1$). The spectrum can be described as optically thick comptonization of a disc blackbody seed spectrum of temperature $k T_0\approx 1.4\,$keV, electron temperature $k T_{\rm e}\approx 10\,$keV and $\tau_{\rm es}\approx 6$, with a corresponding reprocessed component (model C in Table~\ref{tab:may30spec}). An additional soft component is also required: its temperature is rather low: $k T \approx 0.28\,$keV and its amplitude corresponds to inner radius of the emitting disc $\sim 60\,R_{\rm g}$ (for the {\tt diskbb} model). Counts in the first channel are much above the model contribution suggesting more complicated modelling is necessary. This channel was ignored in spectral fitting. \label{fig:may30spec}} \end{figure} \subsection{Analytical comptonization models} A cutoff power law is not necessarily a good approximation to a comptonized spectrum. We first use the {\tt thComp} model as an analytic description of the spectrum, including its self--consistently calculated reprocessed spectrum. We also include a separate Fe K$\alpha$ line from fluorescence from distant material. The presence of this was inferred by Oosterbroek et al.\ (1996) from timing analysis of the May 30th data. Since it is from distant material we fix its energy at 6.4 keV, and assume that it is narrow. We also include a narrow K$\beta$ line at 7.05 keV, tied to 11 per cent of the intensity of the narrow K$\alpha$ line. The best fit {\tt thComp} model has $\chi^2=37.3/27\,$dof, with low electron temperature, $k T_{\rm e} = 9.2\pm 0.4\,$keV, rather large optical depth, $\tau_{\rm es} = 6.73\pm 0.07$, but requires that the seed photon blackbody temperature is very much higher than that of the observed soft flux, with $T_0 = 0.83\pm 0.04\,$keV. Fixing the seed photons to the temperature of the observed soft excess gives a very much poorer fit as it is unable to reproduce the curvature seen in the hard spectrum at low energies. This best fit model also requires a strongly ionized and smeared reprocessed component, with amplitude $\Omega_{\rm r}\approx 0.13$ (Model B in Table~\ref{tab:may30spec}) A significantly better fit can be obtained assuming a broader spectral distribution of the seed photons for comptonization. We have constructed a version of the {\tt thComp} model in which the input soft photon spectrum is a disc blackbody, rather than a simple blackbody. This model (model C in Table~\ref{tab:may30spec}), with its corresponding reprocessed component, can now adequately describe the data, with best fit $\chi^2=25.7/27\,$dof (i.e. $\Delta\chi^2=11.6$ compared to previous case). The comptonizing cloud is rather similar to that derived previously. It is optically thick, $\tau_{\rm es}= 6.49\pm 0.07$ and has temperature of $k T_{\rm e} = 9.6^{+0.7}_{-0.5}\,$keV. The amplitude of the reprocessed component is much smaller than 1, $\Omega_{\rm r} = 0.17^{+0.16}_{-0.05}$, but the reflector properties are nevertheless well constrained. It has to be strongly ionized ($\xi\approx 6000$, so H-like iron ions dominate) and strongly smeared, $R_{\rm in}=6\,R_{\rm g}$. The narrow gaussian at 6.4 keV is still required, its equivalent width is $\sim 70\,$eV. The best fit model spectrum is plotted in Figure~\ref{fig:may30spec}. An almost equally good fit ($\chi^2=29/27$) can be obtained by replacing the gaussian lines by an unsmeared, neutral reflected spectrum, so it is not possible to distinguish from the spectrum whether the distant material inferred from the timing analysis (Oosterbroek et al.\ 1996) is optically thick or thin for electron scattering. Hard lower limit on $N_{\rm H}$ equal to the likely interstellar value of $5\times 10^{21}\,{\rm cm^{-2}}$ was imposed in the above fits. Allowing $N_{\rm H}$ to be a free parameter in model C we actually find that the best fit value is 0, although the fit is now better by only $\Delta\chi^2=2.6$ (F-test significance $\approx 90$ per cent). This may indicate possible complexity of the soft component with further evidence for it coming from the observed excess of counts in the first channel (Figure~\ref{fig:may30spec}). We note however that the excess cannot be solely the result of (unlikely) changes of $N_{\rm H}$ during the integration interval, since the excess remains (although reduced) even if $N_{\rm H}$ is fixed at 0. \subsection{Monte Carlo comptonization models} The analytical solutions for comptonization Green's functions of Sunyaev \& Titarchuk (1980), Lightman \& Zdziarski (1987) and Titarchuk (1994) are not accurate when the seed photon energy is close to the plasma temperature, even in the diffusion approximation. Discrepancies also appear close to the plasma temperature (irrespectively of the seed photons energy) as the models are usually not able to correctly reproduce the shape of the high energy cutoff/Wien peak (see e.g.\ Fig.~2 in Titarchuk 1994). Since in our case the considered energy range is close to both the seed photons energy and plasma temperature, we can expect differences between analytic and simulated spectra. \begin{figure} \epsfxsize = 0.45\textwidth \epsfbox[30 300 550 690]{ext_specs.ps} \caption{Comptonization spectra in the geometry where the source of seed photons is external to the comptonizing cloud {\it cannot\/} explain the unabsorbed spectrum of GS~2023+338. Solid and dashed curves are best fit models C and 0 (Table~\ref{tab:may30spec}), respectively. The seed photons spectrum was assumed a disc blackbody of temperature $T_0=0.28\,$keV, as observed. Spectrum 0 is for $k T_{\rm e}=11\,$keV and $\tau_{\rm es}=6$, i.e.\ the same parameters as model C with central source, but with the external illumination the spectrum is much softer. Spectra {\it a}, {\it b} and {\it c\/} are for $k T_{\rm e}=11,\ 9,\ 5$ keV and $\tau_{\rm es}=8,\ 9,\ 15$, respectively, and are normalized so that they do not exceed the observed spectrum. \label{fig:ext_spec}} \end{figure} However, we demonstrate in Appendix A that for these parameters it is possible to approximate 'exact' spectra (obtained by Monte Carlo simulations) to better than 1 per cent by the analytical models. The results obtained in previous section are therefore vindicated even though we are not able to do proper fitting of Monte Carlo spectra to the data. The spectral curvature observed in the hard component is not an artifact of the deficiencies of analytic comptonization models. The Monte Carlo spectra also require that the seed photons are at much higher temperatures than those from the observed soft excess. The analytic and Monte Carlo models compared here both assume that the seed photons are distributed within the comptonizing cloud. However, this is not the case if the hard X--rays form a central source, surrounded by the accretion disc (see e.g.\ the review by Poutanen 1998 for observational pointers to such a geometry). We computed Monte Carlo spectra assuming a central, spherical plasma cloud of radial optical thickness $\tau_{\rm es}$ and electron temperature $k T_{\rm e}$, surrounded by a geometrically thin disc. The disc is a source of soft photons with a multi-temperature (disc) blackbody spectral distribution parameterized by the temperature at the inner edge, $T_0$. These soft photons then form both the observed soft excess spectrum and act as the {\it external\/} illuminating seed photons for the comptonizing cloud. Figure~\ref{fig:ext_spec} shows examples of inverse Compton spectra obtained in such a geometry. It is not possible to generate a hard and broad enough spectrum in such a geometry, given the constraints on the electron temperature from the observed high energy roll-over and the temperature of the observed soft component. To produce a sufficiently hard spectrum, the optical depth of the cloud has to be large enough for the comptonized spectrum to resemble a Wien peak, which is then rather narrower than the observed spectrum. Thus it seems inescapable that the comptonized spectrum seen is {\it not\/} produced by scattering the observed soft photons, but rather has as its seed photons a rather higher temperature component which cannot be seen directly (probably because it is embedded in the optically thick comptonizing cloud). \subsection{Comparison with Nova Muscae 1991} \label{sec:compar} The spectral and timing behaviour of GS~2023+338 were quite different from those of Nova Muscae 1991 at the peak of its outburst. Takizawa et al.\ (1997) examined {\it Ginga\/} data of Nova Muscae 1991 covering first month of its outburst, when the source luminosity was above $\sim 0.5L_{\rm Edd}$. We computed positions of the peak spectrum of {GS~2023+338}\ on their colour--colour and colour--count rate diagrams (their Figure~9bc). The positions are well outside their diagrams, in the directions indicating that the spectrum of {GS~2023+338}\ is much harder than any of the spectra of Nova Muscae below $\sim 10\,$keV, but it is softer than those above this energy. We illustrate this point in Figure~\ref{fig:gs_nm}, where we plot the spectra of both objects: two spectra of Nova Muscae are plotted: one obtained on 11 Jan ($L\approx 0.5 L_{\rm Edd}$) and the other on 16 Jan, which has $L\approx L_{\rm Edd}$ (number 1 and 6 in Takizawa et al.\ 1997; where the outburst of Nova Muscae was first detected on 8 Jan). The latter spectrum has similar luminosity to the peak spectrum of GS~2023+338, yet is very different in shape. For Nova Muscae the soft component is much hotter ($T_0=0.87\pm 0.03\,$keV) and dominates the luminosity, while the power law is very soft, $\Gamma = 2.76^{+0.08}_{-0.22}$. Even on Jan 11, where the spectrum of Nova Muscae has a strong hard component, its spectrum is much softer than that seen in GS~2023+338 at its peak luminosity. The peak GS~2023+338 spectrum looks similar to the low state spectra seen in both Nova Muscae, GS~2023+338, and other GBH at $L\approx 0.05 L_{\rm Edd}$ except for its roll-over at $\sim 20$ keV, and much stronger soft component. \begin{figure*} \epsfysize = 7 cm \epsfbox[18 460 590 690]{gs_nm_comp.ps} \caption{Comparison of energy spectra of {GS~2023+338}\ and Nova Muscae 1991 at peaks of their outbursts. The spectra (Jan 16 of Nova Muscae and the {GS~2023+338}\ one) are very different even though sources' luminosities (in Eddington units) are very similar and very close to 1. On Jan 11, when $L \approx 0.5\,L_{\rm Edd}$, the soft component cannot be described by a (disc) blackbody -- an additional power law tail is required. The strength of the soft component was comparable to that of the hard component. Further increase in $\dot m$ results in weakening of the hard component (its slope increases), and strengthening of the soft component which now {\it can\/} be described by the disc blackbody model. \label{fig:gs_nm}} \end{figure*} The power spectral density (PSD) of {GS~2023+338}\ during the first 200~s of the $i-1$ interval (excluding the dramatic intensity drop; Figure~\ref{fig:i-1psd}) has intermediate properties between PSDs seen for the two high luminosity spectra of Nova Muscae shown in Figure~\ref{fig:gs_nm} (cf. Figure~2 in Takizawa et al.\ 1997, panels 1 and 6). Its shape is close to a power law, which in Nova Muscae seems to correlate with energy spectra being dominated by a thermal component. Its amplitude is however larger, comparable to the amplitude of the (flat-topped) PSD obtained when the energy spectra show strong comptonized power law. This PSD is nothing like that obtained for GBH low state spectra, where the normalized variability amplitude is much larger, at $\sim 0.1$ at 1~Hz. So even though the energy spectrum bears a (rather small) resemblance to the low state spectrum, its variability behaviour clearly shows that it is very different. \begin{figure} \epsfxsize = 0.45\textwidth \epsfbox[30 200 550 690]{psd.ps} \caption{Comparison of power spectra of Nova Muscae 1991 and GS~2023+338 at peaks of their outbursts. The same data as for energy spectra (Figure~\ref{fig:gs_nm}) are used. Only data between 1--5 keV are used for the 16.\ Jan PSD of Nova Muscae. \label{fig:i-1psd}} \end{figure} Both the spectrum and PSD of GS~2023+338 at its peak luminosity look very different from any known spectral state of GBH. \subsection{Discussion} The spectrum and variability of GS~2023+338 at its peak luminosity show that this is very different from any known GBH spectral state. The soft component is at a very low temperature $\sim 0.3$~keV -- much lower than expected for a disc which extends down to 3 Schwarzschild radii, accreting at close to (or more probably somewhat above) Eddington limit. The hard component looks like a comptonized spectrum where the electrons have a fairly low temperature since the high energy roll-over can clearly be seen. The parameters of the comptonizing cloud are then fairly well constrained. It has to be optically thick, $\tau=5-6$, with electron temperature $\le 10$ keV. These electrons scatter an internal source of soft seed photons whose spectral distribution is broader than a blackbody, with maximum temperature $\ge 1$ keV. These are {\it not\/} the same as the much lower temperature ($\sim 0.3$~keV) observed soft excess. If the seed photons are from the accretion disc then the comptonizing cloud is probably an optically thick corona overlying the accretion disc. This then leads us to a picture where the inner disc is covered by the optically thick comptonizing layers. A rather large extent of the cloud is suggested by simple energetic arguments: since the comptonized spectrum carries $\sim 50$ per cent of the total luminosity, its radius can be expected to be $\sim 25\, R_{\rm g}$, if pure radial stratification of the flow is assumed. We see the normal disc spectrum only from larger radii, which is why the observed soft excess temperature is so much smaller than the $\sim 1$ keV expected from a $12 {\rm M}_{\odot}$ black hole accreting at Eddington rates. The only problem with this picture is that the detected reprocessed spectrum is strongly smeared, to the extent expected if it were reflecting from a disc which extended all the way down to $6R_{\rm g}$. However, the constraints from relativistic smearing can be alleviated if an additional comptonization of the reprocessed component is postulated. Introducing purely phenomenological extension of our model, where the reprocessed component is additionally comptonized in a plasma of temperature $k T_{\rm r}$ and optical depth $\tau_{\rm r}$ (using the Green's function of Titarchuk 1994 in disc geometry), we obtain a good fit, with $\chi^2=19.3/25\,$dof, and the additional plasma parameters $k T_{\rm r} \sim 3.7 $keV and $\tau_{\rm r} \sim 2.5 $. The reflection amplitude is now $f\sim 0.7$ and the inner disc radius is now unconstrained. We note that our implementation of the additional comptonization does not correspond to the effect recently emphasized by Ross, Fabian \& Young (1998). They point out that the reflected X-rays below $\sim 10$ keV will be comptonized when diffusing through strongly ionized outer layers of the reflecting disc, an effect that is not accounted for in {\tt relrepr}\ model (more accurately, this additional comptonization is included in the line profile computations but not in the continuum). While this effect can also be important in our case, as the reflector is indeed strongly ionized, we impose the additional comptonization on the entire reprocessed component, thus changing its overall shape, rather than only below 10 keV. \section{Spectra influenced by strong photo--electric absorption} \label{sec:heavyabs} \subsection{The initial variability} After the steady 200 seconds, the flux from {GS~2023+338}\ dropped dramatically. Is this intrinsic variability, or is it caused solely by photo--electric absorption? As a first example we show a 6--seconds averaged spectrum beginning at $t = t_0 + 329\,$s (Figure~\ref{fig:may30_lc}, \ref{fig:colcol}). The spectrum is plotted in Figure~\ref{fig:s329_abs}, together with the unabsorbed spectrum discussed in Section~\ref{sec:may30}. Plainly they are rather similar at the highest energies, but with a factor of $\sim 3-10$ deficit of counts below 15 keV in the later spectrum. This deficit is far too gradual a function of energy to be caused by complete covering by neutral material, so we first use partial covering by heavy absorption, {\tt thabs}, on a good description of the unabsorbed spectrum (model C in Section~\ref{sec:may30}: disc blackbody and the {\tt thComp} comptonization model with consistent reprocessing). All the unabsorbed spectral parameters are frozen, except for its overall normalization. This gives a very poor fit to the data, with $\chi^2=2028/36$, but including an additional neutral absorber gives a much improved fit with $\chi^2=634/35$, where the absorption parameters are: complete covering with $N_{\rm H}\sim 1.6\times 10^{22}$ cm$^{-2}$, and partial covering with $N_{\rm H}\sim 3\times 10^{24}$ cm$^{-2}$ of 65 per cent of the source. However, the fit underestimates the spectrum in the iron line region and at energies beyond 10 keV. Allowing the additional gaussians to be free gives $\chi^2=130/34$, while including the amount of reflection in the fit gives $\chi^2=77/33$, for an reflected fraction of $\Omega_{\rm r}\sim 0.7$ and line EW of $\sim 290$ eV (compare with $\Omega_{\rm r}\sim 0.17$ and EW $\sim 70$ eV in the unabsorbed spectrum). The complex neutral absorption then has $N_{\rm H} \sim 10^{22}$ cm$^{-2}$ for complete covering and $N_{\rm H}\sim 2 \times 10^{24}$ cm$^{-2}$ covering 65 per of the source. \begin{figure} \epsfxsize = 0.45\textwidth \epsfbox[30 180 550 690]{may30_abs_spec.ps} \caption{Source spectrum (upper panel) on 30 May, 4:36:42 -- 4:36:48, ($\Delta t=6$ sec; 329 sec after the beginning of our data), when the flux dropped by a factor of 3 after an initial, very stable level (cf.\ the light-curve in Figure~\ref{fig:may30_lc}). The spectrum can be described as strongly absorbed spectrum used for the first 200 sec (Figure~\ref{fig:may30spec}; Section~\ref{sec:may30}). The hydrogen column density is $N_{\rm H}\approx 3.5\times 10^{24}\,{\rm cm^{-2}}$, ionization parameter of the absorber $\xi_{\rm abs}\sim 100$ and the covering fraction $\sim 0.62$. The uppermost dashed curve shows the unabsorbed hard component. Lower panel shows the ratio of data to best fit model. The fit is formally unacceptable, $\chi^2=79/33\,$dof, but the residuals do not exceed 5 per cent. They suggest that the absorption cannot be modelled in a single--zone approximation used here. \label{fig:s329_abs}} \end{figure} The apparent changes to the reflection and line emission can be caused by our use of {\tt thabs} to describe the very heavy absorption. At such high columns, Compton down-scattering can distort the shape of the absorbed spectrum. We replace the analytic approximation by our Monte Carlo transfer code to model the heavy partial absorption. With only the line emission allowed to be free we obtain $\chi^2=90/34$, with absorption parameters $N_{\rm H}\sim 1.4\times 10^{22}$ cm$^{-2}$, and partial covering with $N_{\rm H}\sim 3\times 10^{24}$ cm$^{-2}$ of 62 per cent of the source. For this model, the normalization of the underlying spectrum is less than 5 per cent different to that of the unabsorbed spectrum except that the addition (narrow gaussian) iron line intensity has increased by a factor of $\sim 3$. The fit can be improved somewhat by allowing the heavy partial absorber to be ionized. This is the best fit spectrum shown in Figure~\ref{fig:s329_abs}. The column density is $N_{\rm H}=3.5\times 10^{24}\,{\rm cm^{-2}}$ i.e.\ $\tau_{\rm es}\approx 2.8$, $\xi_{\rm abs}\sim 100$ and the covering fraction $\sim 0.6$. However, even this gives $\chi^2\sim 79/33$, i.e.\ the fit is formally unacceptable with residuals at the 5 per cent level. The fit cannot be improved by small changes to the parameters of the primary (comptonized) spectrum or the soft component. Thus even though the overall spectrum is fairly well explained by the absorption hypothesis, the presence of significant residuals points towards complexity of the absorber. Since the data are an average over a long enough time for the absorber to change, a model involving a range of column densities may be required, rather than a single--zone approximation used here. \subsection{All variability} We use the colour--colour diagram as a guide to the variability. On May 30th the source showed dramatic spectral variability (Figure~\ref{fig:colcol}), and this continued during the first month of the decline from outburst (Figure~\ref{fig:heavyabs}). The unobscured spectra are easy to spot on such diagrams, since they fall at the leftmost end of a distinct horizontal track in the colour--colour plots (see also Oosterbroek et al.\ 1997). The unusual spectrum seen at the unobscured peak outburst luminosity of {GS~2023+338}\ is {\it not\/} the same as the unobscured spectra seen several days later (see Paper I). The unobscured spectra at the end of the absorption track from June 3rd onwards are fairly typical low/hard state spectra (see Paper I), and have inferred luminosities of $\le 5$ per cent of $L_{\rm Edd}$. Hence there must be some intrinsic spectral and intensity variability as well as the photo--electric absorption. We will not attempt to construct a full time history of the absorber but will limit our efforts to identifying the spectral components present when the source was in various locations on the colour--colour diagram. Because of the dramatic variability (often on 1 sec time scale), the time intervals over which we summed the spectra are a compromise between requirements of a stable position on the colour--colour diagram and photon statistics. We generally assume that the intrinsic spectrum consists of three components: a soft thermal component which may or may not contribute the seed photons for Compton upscattering to make the hard power law, the comptonized power law and its corresponding reprocessed component. The absorbing material, as well as distorting the intrinsic spectrum, may produce fluorescent line emission, or a reprocessed component of its own. In the May 30th $i-1$ interval we have shown that the spectral and intensity variability is consistent with complex, heavy absorption. The $i-2$ interval shows similar behaviour to the $i-1$ interval on the colour--colour plot (Figure~\ref{fig:colcol}), with strong variability in the soft colour but little change in the hard colour. Thus it seems likely that most of the variability during this interval is also caused by a changing column of the heavy absorption which covers 60--70 per cent of the source, while the intrinsic spectrum remains more or less constant. The variability in the $i-3$ interval is apparently different, starting off in a very low intensity state which is much harder at high energies (approximate position in the diagram $[1.4,2.5]$), which then rises to $[2,3]$, and then rejoins the $i-1$ and $i-2$ absorption track at $\sim [3,1.5]$. However, even here the spectrum is roughly consistent with that of the peak luminosity state (in both shape and intensity), with the heavy absorption changing from a covering fraction of $\ge 90$ per cent at the start to $\sim 70$ per cent as seen in the $i-1$ and $i-2$ intervals at the end. However, the $i-4$ spectrum is completely different. It cannot be fit by any form of absorption of the peak luminosity spectrum. There is a distinct hard tail, showing that there is indeed some heavy photo--electric absorption, but the low energy spectrum is much softer than that of the peak spectrum (as shown by it having a smaller hardness ratio at low energies). The intrinsic spectrum has changed! We can model this spectrum using a soft component whose temperature is now $\sim 1$ keV. The hard X--ray spectrum is then compatible with these being the seed photons for Compton upscattering, forming a fairly steep power law with $\Gamma\sim 2$, and where the rollover from the electron temperature is not detectable. This is then partially covered by heavy photo--electric absorption. We will discuss the nature of the intrinsic spectrum in more detail in the next section. Here we will merely note its resemblance to the classic high state spectrum (high temperature, steep power law component). The total bolometric unobscured luminosity is then derived to be $\sim 7\times 10^{-8}$ ergs cm$^{-2}$, equivalent to $ 0.07 L_{\rm Edd}$. None of the data taken after this time show a significant high energy rollover, and all have a hard component which can be described by comptonization of the observed soft photons. We use this interval as the starting point (data taken after May 30th 06:30), and plot the colour--colour diagram for all the data up to June 5th (see Figure~\ref{fig:heavyabs}). We have extracted spectra from a number of regions on the colour--colour diagram, and modelled them to try and deconvolve the intrinsic spectrum from the effects of absorption. During $i-4$ and $i-5$ the source changes mainly in hard colour, making a vertical track (track A) in the colour--colour diagram (Figure~\ref{fig:heavyabs}). This track continues in $i-6$ and $i-7$, but here the spectral variability becomes faster and more chaotic, and includes changes in soft colour as well. Good fits are obtained to spectra {\it i}--{\it h}--{\it g} along track A with a model in which the soft {\tt diskbb} component has little absorption, while the hard component (Compton scattered power law tail) is more strongly absorbed. In addition, this hard power law illuminates cool, neutral material, giving a reflected spectrum which is even more strongly absorbed, and whose normalisation is strongly enhanced with respect to the observed hard component. Motion from the bottom of track 'A' to the top corresponds to increasing the enhancement of this reflected component from about a factor 7 at spectrum {\it i\/} to about a factor 100 at spectrum {\it g\/}. The soft spectrum is consistent with remaining fairly constant throughout this track (see Oosterbroek et al.\ 1997), while the hard component shows the normal broad band variability (see below). \begin{figure*} \epsfxsize = \textwidth \epsfbox[18 144 592 718]{absspec.ps} \caption{ The colour--colour diagram (cf.\ Figure~\ref{fig:colcol}) and examples of spectra influenced by heavy photo--absorption. Solid squares are the source's positions on May 30th, 6:30--10:00 (i.e.\ only part of the May 30th data is represented here); small open circles: June 3rd -- 5th. The labelled circles mark approximate locations where the spectra were extracted. The spectra are assumed to consist of a soft component modelled as a disc blackbody, absorbed primary power law and its corresponding (but independently absorbed) reprocessed component, together with gaussian lines at 6.4 keV and 7.05 keV (Fe fluorescence), as labelled. Track 'A' is consistent with both the power law and the reprocessed component being absorbed below $\sim 4$ keV. Movement downwards along the track corresponds to gradual disappearance of the reprocessed component. Horizontal movement corresponds to changes of the normalization of the primary power law. Track 'B' begins with unabsorbed, low state spectra (position: 1.5, 1.2) and increasing soft colour corresponds to stronger absorption, influencing the primary power law and its reprocessed component. \label{fig:heavyabs}} \end{figure*} The total unobscured luminosity at the top of track A is then $\sim 0.5 L_{\rm Edd}$ so the inferred luminosity increases by a factor $\sim 10$ from the bottom to top of track A. Yet it seems rather unlikely that this change is actually intrinsic, since the soft component shows very little change in temperature or normalization. It is more likely that the source remains fairly constant at $0.5L_{\rm Edd}$, but that some (variable) amount of the source is completely obscured by optically thick material, and that we see only the intrinsic source spectrum via scattering. The scattering origin for the observed soft component would be in accord with its small normalization, which is only $\sim 1$ per cent of that expected, if the component indeed is a disc emission from $6\,R_{\rm g}$ onwards. We note here the strong constraints on the scattering scenario for the hard component coming from its observed broad band variability. The high amplitude of the PSD of the hard component ($E>5\,$keV) during the $i-5$ interval (Figure~\ref{fig:may30_psd}) means that the scatterer could not have been located farther than $\sim 0.1$ lsec $\sim 10^3\,R_{\rm g}$ (see also next Section). One geometrical scenario in which to understand these results is that there is a thick disc which blocks our direct view of the source, so that we see only the scattered fraction of the intrinsic radiation. The obscuring wall is axi--symmetric, so there is a strong reflected component of the intrinsic (rather than scattered) emission from the opposite wall. As the disc subsides then we see progressively more of the reflector, so the strength of the hard component increases, giving the variability seen as track A. Eventually the optically thick material goes below our line of sight, allowing the intrinsic hard spectrum to be seen. Differing amounts of fairly heavy absorption on this intrinsic hard spectrum then cause the variability in soft and hard colour defined by spectra {\it g}--{\it f}--{\it e}, but the intrinsic spectrum is consistent with remaining approximately the same in both shape and intensity. Thus despite the dramatic spectral and intensity variability, the data on May 30th are compatible with very little intrinsic variability. At the start of the observation the source was accreting at (super) Eddington luminosities, with a spectrum unlike any of the known (sub Eddington) spectral states. It then declined by only a factor of $\sim 2$, and made a state transition to the high (or very high) state, and remained at this level for the rest of the observation. All the rest of the variability seen is connected to the absorption. After the May 30th observation the next data were not taken until 48 hours later, mid-morning of June 1st. These were taken with satellite pointing position which was offset by $\sim 0.8^\circ$ from the source, so cannot be used for detailed spectral analysis. However, on the colour--colour plot they span the region around spectrum {\it d}. The next good data are taken 12 hours later, on the early morning of June 2nd, where the source spectrum again looks very similar to that of spectrum {\it d}. At first sight this looks like a continuation of the {\it g}--{\it f}--{\it e} track, but fitting of the data shows distinct differences. Firstly the overall intensity is rather lower, with the hard component being a factor 10 weaker than in spectrum {\it e}. The soft emission is also at much lower temperatures. These, together with the good data taken on June 3-5th form a distinct track in Figure~\ref{fig:heavyabs} along spectra {\it d}--{\it c}--{\it b}--{\it a} (track B) which is {\it not\/} simply connected to track A. Again there must have been some intrinsic spectral change, since the unabsorbed spectra at the end of this track show a classic low state spectral form (Paper I). Track B can be reproduced by cold partial covering of the Jun 3rd unabsorbed low state spectrum, fixing all the parameters to be the same as in the unabsorbed data (see Paper I). In this model the overall normalization of the June 3rd spectrum (which gives a total bolometric flux of $\sim 4\times 10^{-8}$ ergs cm$^{-2}$ s$^{-1}$ or $0.04L_{\rm Edd}$) changes by less than 25 per cent while the partial covering absorption parameters change from $3 \rightarrow 7 \rightarrow 12 \times 10^{22}$ cm$^{-2}$ covering $\sim 92$ per cent of the source going from spectra {\it a}$\rightarrow${\it b}$\rightarrow${\it c}. This confirms the suggestion by Oosterbroek et al.\ (1997) that the track corresponds roughly to increasing $N_{\rm H}$. We have however used here a much more realistic intrinsic spectrum -- the unabsorbed spectrum of the June 3rd data set (Paper I). The turning point of track 'B' corresponds to the absorption column of $\approx 1.5\times 10^{23}\,{\rm cm^{-2}}$. Further increasing $N_{\rm H}$ leads to a decrease of the soft colour since it decreases the counts in the 3--5 keV band, whilst the counts in the 1--3 keV band -- determined by the soft component -- are constant. Thus spectrum {\it d\/} can be fit into the above pattern, with a column of $\sim 31\times 10^{22}$ covering 92 per cent of the source. \section{Has the source ever made a transition to a soft state?} \label{sec:soft} As we point out in Section~\ref{sec:heavyabs} some of the spectra during May 30th observation show a thermal component of temperature $\sim 0.8\,$keV (for {\tt diskbb} model) and power law tails with $\Gamma \sim 2$, typical for high/soft states of GBH. To further characterize the source in those time periods we examined more closely some of the energy spectra as well as the source variability. \begin{figure*} \epsfysize = 7 cm \epsfbox[18 420 600 690]{psd_may30.ps} \caption{Power spectra of X-ray emission obtained on two occasions: panel ({\it a}\/) shows PSD on 30th May (time interval $i-5$, 07:26 to 07:48; see Figure~\ref{fig:may30_lc}), panel ({\it b}\/) shows PSD on 20th June when the source was in the usual low/hard state (Paper I). Labels in panel {\it a\/ } show energy bands in keV. The thin dotted line represents PSD from panel {\it b\/} for direct comparison. On 30th May the $1-3$ keV component did not vary on time-scales shorter than $\sim 100$ sec, but the amplitude of variability increased with energy. In all three spectra an enhanced variability on time-scales longer than $\sim 100$ sec was observed. This can be attributed to variable photo-electric absorption. On 20th June (panel {\it b\/}) the variability was independent of energy. \label{fig:may30_psd}} \end{figure*} \subsection{Timing and variability} \label{sec:soft_timing} Power spectra of {GS~2023+338}\ were already examined by Oosterbroek et al.\ (1997). Their Fig.~8 (panel 1A) shows PSD on May 30 exhibiting strongly reduced power above $\sim 1$ sec, similarly to what is usually observed in high state (van der Klis 1995). In Figure~\ref{fig:may30_psd}a we plot PSD for data obtained between 07:26 and 07:48 ($i-5$; see Figure~\ref{fig:may30_lc}), separately for three energy bands: $1-3\,$keV, $3-5$ keV and $5-30\,$keV. The $1-3$ keV band variability is consistent with pure Poisson noise on time-scales $\delta t \la 100$ sec whilst the harder X-rays are highly variable with the amplitude of variability increasing with energy up to $\approx 5$ keV. For comparison Figure~\ref{fig:may30_psd}b shows PSD for data taken on 20th June (cf.\ Miyamoto et al.\ 1992), when the source was in the usual low/hard state (Paper I). Here the amplitude of variability does not depend on energy. This analysis confirms that the energy spectrum consists of two components, a constant, soft component, and a strongly variable hard component, in accord with spectral decomposition shown in Figure~\ref{fig:soft_spec}. A characteristic feature of the PSD is the increase of power for $\delta t>100$ sec. Since even the soft component varies on that time-scale, the increase is most likely due to slowly variable absorption. The PSD of the hard component (5 -- 20 keV) on May 30th is rather steeper than that on 20 June for $\delta t \la 10$ sec (frequency $f \ga 0.1$ Hz). The loss of power could be due to reflection/scattering in an extended medium whose presence is suggested both by the energy spectra analyzed in the next section and by the previously discussed increase of PSD on long time-scales. Alternatively, it could be an intrinsic feature of the primary emission. If so, however, then the steepening of PSD in putative soft state would be opposite to what is shown by e.g.\ Cyg X-1 (van der Klis 1995 and Cui et al.\ 1997). \subsection{Energy spectra} \begin{figure*} \epsfysize = 7 cm \epsfbox[18 440 590 690]{int4_eufs.ps} \caption{Examples of spectral modelling of the $i-4$ spectrum, when the source was heavily absorbed. Model {\it a\/} is purely phenomenological, comprising two power law spectra absorbed by two (partial) absorbers and an overall absorption. Model {\it b\/} consists of an unabsorbed disc blackbody with $k T\approx 1$ keV, and a power law ($\Gamma\approx 2$) with two absorbers ($N_{\rm H,1}\approx 24$, $f_1=1$; $N_{\rm H,2}\approx 300$, $f_2=0.86$). Model {\it c\/} has similar disc blackbody, a power law with $\Gamma=2.3$ absorbed by $N_{\rm H}=27$ and $f=1$, and an (enhanced) reprocessed component absorbed by $N_{\rm H}=180$ and $f=1$ (all values of $N_{\rm H}$ are given in units of $10^{22}\,{\rm cm^{-2}}$). Additional narrow gaussian lines at 6.4 and 7.05 keV are included to account for fluorescence in the absorbers. Values of $\chi^2$ are, respectively: $21.2/22\,$dof, $20.6/23\,$dof and $17.4/22\,$dof. Only models {\it b\/} and {\it c\/} are compatible with PSD shown in Figure~\ref{fig:may30_psd}a, i.e.\ the variability increasing with energy up to $\sim 5\,$keV. Parameters of those models are typical for high/soft state of SXT. \label{fig:soft_spec}} \end{figure*} We have extracted two spectra of {GS~2023+338}\ from time intervals: $i-4$ at 6:43:45 -- 6:47:20 and $i-5$ at 7:26:45 -- 7:49:00 (see light-curve in Figure~\ref{fig:may30_lc}, and position on the colour--colour diagram, Figure~\ref{fig:colcol}). First, we tested the hypothesis that the spectra can be described assuming (possibly non-uniform) absorption acting on a typical hard state spectrum. To this end we assumed the June 3rd (see Table~2 in Paper I) spectrum and allowed the three components (disc blackbody, power law and reflection) to be absorbed. This model fails in both cases even though we allowed the three absorbers to be different. The best fits have $\chi^2=1998/22\,{\rm dof}$ and $\chi^2=571/22\,{\rm dof}$. Adding to the model two narrow gaussians at 6.4 and 7.05 keV (to account for iron fluorescence from the absorber) improves the fits significantly but they are still unacceptable: $\chi^2=1220/21\,{\rm dof}$ and $\chi^2=56.3/21\,{\rm dof}$. \begin{table*} \caption{Model fitting of the soft state spectra. \label{tab:soft_spec}} \begin{tabular}{lcccccccr} data & $k T_{\rm soft}$ (keV) & photon index & $N_{\rm H,1}\,{\rm cm^{-2}}$ & $\Omega_{\rm r}$ & $\xi$ & $N_{\rm H,2}$ & EW (eV) & $\chi^2/$dof \\ \hline $i-4$ & $1.07^{+0.04 }_{-0.11}$ & $2.24\pm 0.16$ & $(25^{+5}_{-8})\times 10^{22}$ & $7.4^{+2.2}_{-1.1}$ & $100^{+400}_{-100}$ & $(180^{+30}_{-20})\times 10^{22}$ & $430^{+70}_{-110}$ & 17.4/21 \\ $i-5$ & $0.98^{+0.03 }_{-0.06}$ & $1.90^{+0.13}_{-0.11} $ & $(24\pm 5)\times 10^{22}$ & $6.3^{+1.0}_{-0.8}$ & $25^{+40}_{-24.8}$ & $(130^{+27}_{-20})\times 10^{22}$ & $375^{+75}_{-60}$ & 15.8/21 \\ \hline \end{tabular} \medskip Model: abs*( disc blackbody + abs1*(power law) + abs2*({\tt relrepr}\ ) + gaussian + gaussian) \end{table*} The simplest phenomenological description of the two spectra consists of two power law components with different absorbing columns, $N_{\rm H}$, and covering fractions, $f$, with the two narrow gaussian lines and an overall absorption. One power law is rather soft, $\Gamma\sim 3$ and it is absorbed by a column of $N_{\rm H}\sim (6-8)\times 10^{23}\,{\rm cm^{-2}}$ and $f = 0.65-0.9$. The other is harder, $\Gamma\sim 1.2$ while $N_{\rm H}\sim 2\times 10^{24}\, {\rm cm^{-2}}$ and $f=1$. In an attempt to construct a more realistic model we assumed an unabsorbed disc blackbody as the soft component and an absorbed power law, but such a model fails to describe the data. Better fits are obtained when a second, partial absorption is applied to the power law: $\chi^2=20.5/22\,$dof for $i-4$ and $\chi^2 = 27.8/22\,$dof for $i-5$, with the second absorbing column $\sim 3\times 10^{24}\,{\rm cm^{-2}}$. The fit to $i-5$ data can be further improved if the second absorber acts on a different power law component. One realistic candidate for the second power law is the reprocessed component whose presence is also in line with our previous results. Replacing the power law with the (non-smeared) reprocessed component we obtain good fits with $\chi^2=17.4/21\,{\rm dof}$ for $i-4$ and $\chi^2=15.8/21\,$dof for $i-5$. The model parameters are well constrained: the temperature of the soft component is fairly high, $T_0 \sim 1\,$keV, whilst the power law index is $\sim 2$. Both values are indeed as expected in typical high/soft state spectra of GBH. The spectra are plotted in Figure~\ref{fig:soft_spec} and fit results shown in Table~\ref{tab:soft_spec}. \subsection{Discussion} Based solely on properties of the $i-4$ and $i-5$ spectra it is impossible to determine whether the constancy of the soft component is its intrinsic property or it results from the component (or a fraction of it) being scattered off some extended scatterer towards the observer. The lack of variability is intrinsic if the hard component is scattered as well -- as the overall evolution seems to suggest -- since the scattering preserved the variability characteristics of the hard component. An intrinsically weak variability then supports the spectral identification of the soft component as the thermal disc emission since this is known to be weakly variable (van der Klis 1995). The intrinsic luminosity of the soft component is $\approx 9\times 10^{38} \,{\rm erg\,cm^{-2}\,s^{-1}} \approx 0.6\,L_{\rm Edd}$ ($d=3.5\,$kpc), assuming that it indeed is a disc emission from $6\,R_{\rm g}$ onwards. Effective temperature expected at the inner disc edge is then $\approx 1.1\,$keV (Frank, King \& Raine 1992), in surprisingly good agreement with the best fit value (fitting the {\tt diskspec} model with the colour temperature correction 1.5). The energy spectrum of {GS~2023+338}\ during the $i-4 \rightarrow i-7$ time intervals is thus compatible with typical high/soft state spectra of GBH. However, the strong, broad band noise variability of the hard component is not. Examining more closely the {\it Ginga\/} data of Nova Muscae 1991 (see Section~\ref{sec:compar}) we find that in the typical high/soft state the hard component ($E>5\,$keV) shows almost no variability, similarly to the soft component (note that in the {\it Ginga\/} data, counts are dominated by channels at 3--5 keV, so the usually plotted -- summed over all energies -- PSDs are dominated by the soft component). Perhaps a weak intrinsic variability is enhanced by absorption, as suggested by the fact that the PSD amplitude of the soft component was larger later, in $i-7$ than in $i-4$ and $i-5$. \section{Discussion and Conclusions} \subsection{Intrinsic X--ray Spectral Evolution} Much of the evolution of the X--ray spectrum is hidden beneath a veil of complex, heavy absorption, and it is not generally possible to uniquely recover the intrinsic spectrum. Most of the dramatic spectral {\it and intensity\/} variability seen on May 30 is connected with the evolution of the complex absorption rather than with the X--ray source itself. The apparent saturation of the X--ray luminosity at $\sim 10^5$ counts sec$^{-1}$ (see e.g.\ Figure~\ref{fig:may30_lc}) is {\it not\/} due to the source dramatically flaring and then hitting the Eddington limit, but instead can be explained by the source staying fairly constant while our direct line of sight to it is covered (and uncovered) by very optically thick material. The absorption variability is such that there are occasional glimpses of the unobscured source. One such time is at the start of the observation on May 30. Here the observed luminosity is at least 0.6 of the Eddington limit (integrating the derived spectrum to get the bolometric luminosity gives the model dependent number of $\ge 1.6\timesL_{\rm Edd}$), and the spectral shape is not at all like that seen from other transient systems (at any luminosity!). At these high accretion rates we might expect to see a strong soft component from the inner accretion disc at around $\sim 1$ keV, yet the observed strong soft emission is at temperatures around $0.2$ keV, much lower than expected. The hard X--ray spectrum is not a power law. Instead it has a distinct roll-over, indicating relatively low electron temperatures, $\sim 10$ keV, in the hard X--ray emitting plasma. The curvature of this spectrum is rather difficult to match unless the seed photons for this comptonized spectrum are at temperatures of $\sim 1$ keV. Thus it seems most likely that the missing accretion disc photons are hidden under the optically thick comptonizing cloud which produces the hard X--ray spectrum. Slim disc models (including the advection of trapped radiation) of super-Eddington rates may point towards the formation of such a hot central region (Beloborodov 1998). These extreme luminosities are probably accompanied by a strong wind driven from the disc, which is perhaps the origin of some of the dramatic absorption variability. The source then makes a transition to the standard very high or high state spectrum, with the expected strong soft component at $\sim 1$ keV. This transition probably takes place between the $i-3$ and $i-4$ time intervals (see Figure~\ref{fig:may30_lc}), perhaps connected to the source luminosity decreasing from $\sim 1.5\times L_{\rm Edd} $ to slightly below $L_{\rm Edd}$. All the May 30th spectra are consistent with an intrinsic bolometric luminosity close to $L_{\rm Edd}$, although this is highly model dependent due to the obscuration of the source. The source was not observed on May 31, and no good data exist for the observation on June 1st. This is somewhat unfortunate, since on June 2nd the data are consistent with the source having an intrinsic luminosity of $\sim 0.04-0.05 \,L_{\rm Edd}$, and show the standard low/hard spectrum. Somewhere in the two missing days the intrinsic source luminosity decreased by a factor of 10--20! The source then decreases by about a factor of 2 from June 2nd to July 6th, consistent with a standard e--folding decay timescale of 30--40 days. In summary, all of the oddities of {GS~2023+338}\ may perhaps be explained by the source accreting at super-Eddington rates. This would give the unusual spectrum seen at the start of May 30th, and power a strong outflowing wind which caused dramatic absorption variability. As the source declined below Eddington luminosity it showed the standard high state spectrum. It may have further declined steadily (although rapidly) through the high state spectrum to the low/hard state on May 31st--June 1st, or there may have been a more dramatic event, perhaps linked to the observed transient radio emission, in which there was complete disruption of the inner disc. Perhaps this signaled a huge ejection of the accreting matter, so that the accretion rate onto the central object was much reduced, and adding to the complex absorption. \subsection{Disc Evolution} \label{sec:discuss} The most promising explanation of SXT outbursts seems to be the classical disc instability model, modified to include the effects of X-ray irradiation (e.g.\ Cannizzo 1993, van Paradijs 1996, King \& Ritter 1998). The quiescent disc builds up from matter accreted from the companion star until the outburst is triggered. The disc switches into the hot, ionized state, and has the familiar Shakura--Sunyaev structure. The outer disc temperature eventually drops below the H ionization temperature, and a cooling wave propagates inwards, switching the disc back into quiescence. Without X--ray irradiation the outburst lightcurves are the classic drawf nova lightcurves, with a linear decline after outburst. However, if X--ray irradiation is strong enough to keep the outer disc ionised then the cooling wave is supressed. Most of the disc mass can then accreted before the cooling wave can form, giving an exponential decay (King \& Ritter 1998, King 1998). The prevalence of exponential decays in SXT X--ray lightcurves point to the importance of irradiation in these systems (Shahbaz, Charles \& King 1998). {GS~2023+338}\ shows a fast rise, followed by an exponential X--ray decay. The fast rise suggests that the outburst is triggered towards the outer edge of the disc (Smak 1984, Cannizzo 1998), so that all of the disc takes part in the initial heating wave. The exponential decay means that we might expect that irradiation (probably indirect via scattering in a corona: Dubus et al.\ 1999) is important so that most of the disc stays in the outburst state, and so is accreted. Integrating the observed X-ray lightcurve gives an estimate of the accreted mass $\Delta M_{\rm X} \sim 6\times 10^{25}\,$g, assuming a radiative efficiency of 10 per cent. This matches very well with estimates of the mass transferred from the companion star during the inter-outburst time interval $\Delta M_{\rm c} \approx 2\times 10^{26} $g, adopting $\Delta t \approx 32\,$years; (Chen et al.\ 1997) and the mass transfer rate from the companion star $\dot M_{\rm c}\approx 10^{17}\,{\rm g\,s^{-1}}$ (King, Kolb \& Burdieri 1996). However, this disc mass is orders of magnitude smaller than the disc mass that would be built up under the assumption that the outburst is triggered by the surface density of the quiescent disc reaching its critical value everywhere. The orbital period of {GS~2023+338}\ is 6.5 days, so the disc can extend out to radii of $R_{\rm out} \simeq 1.36R_{\rm circ} \approx 0.7 R_{\rm tidal} \approx 1.2\times 10^{12}$ cm (Shahbaz et al.\ 1998). This predicts a huge disc mass of $M_{\rm disc} \sim \rho R_{\rm out}^3 /3 \sim 2 \times 10^{28}$ g ($\rho\sim 3\times 10^{-8}\,{\rm g\,cm^{-3}}$; King \& Ritter 1998). Plainly this shows that the outburst is triggered long before this maximum disc mass is built up. A similar result is found for the other well studied large disc system, GRO J1655-40 (Shahbaz et al.\ 1998). This is in marked contrast the to short period (small disc) SXT's, where the disc mass derived from the maximum quiescenct disc calculations is (roughly) equal to the mass inferred from the X--ray luminosity (Shahbaz et al.\ 1998), and to the integrated mass transfer rate from the companion star (Menou et al.\ 1999). Perhaps a clue to resolving the problem is that the region where the disc solution is unstable in {GS~2023+338}\ is quite distant from $R_{\rm out}$. Solving the vertical disc structure equations (using the code recently described in R\'{o}\.{z}a\'{n}ska et al.\ 1999), we find that the solution is unstable where $T_{\rm eff}=(5 - 7) \times 10^3\,$K (see also e.g.\ Hameury et al.\ 1998). This gives $R_{\rm unst} = (5 - 8) \times 10^{10}\,$cm, i.e.\ the ring is centered at $\approx 0.05\,R_{\rm out}$. In Nova Muscae 1991 ($P_{\rm orb}=10.5\,$h) $R_{\rm unst}$ is rather closer to $R_{\rm out}$: $R_{\rm unst} = (2 - 3) \times 10^{10}\,{\rm cm} = (0.2 - 0.3) R_{\rm out}$. Perhaps the disc beyond a few $R_{\rm unst}$ never builds up to full quiescence, but instead stays on the steady state, cool branch. Whatever the reason, it seems that there are serious deficiencies in our understanding of the structure of large discs. \section*{Acknowledgments} We thank Andrei Beloborodov for helpful discussions on super-Eddington accretion, and Bo\.{z}ena Czerny, James Murray and John Cannizzo on accretion disc instabilities. CD acknowledges support from a PPARC Advanced Fellowship. Work of PTZ was partly supported by grant no.\ 2P03D00410 of the Polish State Committee for Scientific Research.
\section{Introduction} \setcounter{equation}{0} It is none to say what group theory meant (and means) for theoretical physics during this century, in particular the theory of continuous groups or Lie groups ( see e.g. \cite{lie1}, \cite{lie2} and references therein). The properties of their Lie algebras, easier to hand than the groups themselves, define them locally and in fact most part of the textbooks are dedicated to them \cite{lie3}. However depending on the problem at hand, explicit parametrizations of the group manifold become necessary. Many of them are widely known, example of them the $SU(2)$ or more generally the Euler angles of orthogonal groups. Of course that we can always write locally a group element as a product of one-generator exponentials, or simply as the exponential of an arbitrary Lie algebra element. But in most cases theses obvious parametrizations are of few usefulness because they obscure the global properties of the group and generally lead to un-tractable computations. An interesting parametrization is suggested by the Mackey theorem, the well-known coset decomposition: let $G$ be a group and $H$ a subgroup of it. Then for any $g\in G$, \begin{equation}}\newcommand{\ee}{\end{equation} g = k\; h , \;\; k\in G/H\; , \;\; h\in H \label{coset} \ee in a unique way. This leads to the theory of homogeneous spaces (or (left or right) coset spaces) $G/H$, good references on the subject being \cite{hel}, \cite{gil}. Among others physical applications \eqn{coset} is fundamental in the treatment of effective field theories with spontaneously broken symmetries \cite{wein}. But let us assume instead that we have a field theory including maps from ``space-time'' on a group manifold $G$ among its degrees of freedom, and gauge invariant under the adjoint action of a subgroup $H$ of $G$ \begin{equation}}\newcommand{\ee}{\end{equation} g\;\rightarrow\; {}^h g = h\; g\; h^{-1}\;\; ,\;\; h\in H \label{adja} \ee This means that effectively the theory depends on the invariants of the group under the adjoint action of the subgroup. That is, if we were able to write uniquely any element of $G$ as \begin{equation}}\newcommand{\ee}{\end{equation} g \equiv h^{-1}\; \bar g\; h \;\; , h\in H \label{adj} \ee then clearly making a gauge transformation \eqn{adja} identifying the $h$'s the theory will depend only on $\bar g$ that encloses the invariants mentioned above. We would like to remark that at difference of \eqn{coset} there no exists any general theorem assuring the decomposition \eqn{adj}; in fact it is not difficult to find examples where it is not possible to write it. It is the aim of this paper to get the class of local parametrizations of the type \eqn{adj} in a whole set of cases of physical interest. Specific examples where they must be used (and were used in the lower dimensionality cases where parametrizations were available) are the two dimensional gauged Wess-Zumino-Witten-Novikov models \cite{gwzw}, \cite{gaw}. It is worth to say however that the results, being explicit parametrizations of classical groups, are valuable on their own right independently of the applications. \section{The orthogonal groups} \setcounter{equation}{0} We consider in this section the pseudo-orthogonal groups, $G \equiv SO(p,q)$. Its maximal compact subgroup is $H\equiv SO(p)\times SO(q)$. In order to get its decomposition \eqn{adj} we need as a first step to get the \subsection{Reduction of $SO(p+1)$ under $SO(p)$.} We start by writing \cite{gil} \begin{equation}}\newcommand{\ee}{\end{equation} P_{p+1}(\vec u_p , P_p ) = K_p(\vec u_p )\; H(P_p, 1)\label{cosort} \ee where $\;\vec u_p\;$ is a $p$-dimensional real vector, $P_p\in SO(p)$ and generically we will mean \begin{equation}}\newcommand{\ee}{\end{equation} H(P,Q) = \matriz{P}{0}{0}{Q} \ee In what follows the dimensionalities of matrices should be understood from the context when not stated explicitly. The right coset element in $SO(p+1)/SO(p)\sim S^p$ is given by \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} K_{p+1}(\vec u_p ) &=& \left( \matrix{ 1 - (1 + u_{p+1} )^{-1}\; \vec u_p \vec u_p{}^t &\vec u_p\cr -\vec u_p{}^t &u_{p+1}\cr } \right)\cr 1 &=& \vec u_p{}^t \vec u_p + (u_{p+1})^2 \eea Under an adjoint transformation \begin{equation}}\newcommand{\ee}{\end{equation} {}^h P_{p+1} = H(h,1)\; P_{p+1}\; H(h^t ,1)\;\; ,\;\; h\in SO(p) \ee the parameters of $P_{p+1}$ get transformed as: \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} {}^h \vec u_p &=& h \; \vec u_p \cr {}^h P_p &=& h\; P_p \; h^t \label{adjtras1} \eea The procedure will be constructive. Les us pick an arbitrary matrix $\;V_p\in SO(p)\;$ decomposed this time as a left coset w.r.t. $SO(p-1)$ \begin{equation}}\newcommand{\ee}{\end{equation} V_p = H(V_{p-1}, 1) \; K_p(\vec v_{p-1}) \label{parV} \ee and rewrite \eqn{cosort} for any such a $V_p$ with the help of \eqn{adjtras1} as \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} P_{p+1} &=& H(V_p{}^t ,1)\; {\bar P}_{p+1} \; H(V_p , 1)\cr {\bar P}_{p+1} &=& K_{p+1}(V_p\vec u_p )\; H(V_p P_p V_p{}^t ,1)\label{adjort} \eea The general idea to apply here and in the subsequent cases is to fix the whole matrix $V_p$ ($\equiv \vec v_{p-1}, V_{p-1}$ ) in terms of variables of $P_{p+1}$ ($\equiv \vec u_p , P_p$), leaving aside only precisely the invariants together with the matrix $V_p$ as parameters of $P_{p+1}$. Evidently this procedure is equivalent to make a change of variables from the non-invariant parameters in $P_{p+1}$ to $V_p$. Equation \eqn{adjort} suggests to put $\;\vec u_p\;$ in some standard form by a specific choice of (a part of) $V_p$. In fact it is easy to show that the choice \footnote{ As usual we will use the notation $({\check e}_i )_j = \delta_{ij}, (E_{ij})_{kl}=\delta_{ik} \delta_{jl}, A_{ij} \equiv E_{ij} - E_{ji}, S_{ij} \equiv E_{ij} + E_{ji} $.} \begin{equation}}\newcommand{\ee}{\end{equation} \left( \begin{array}{c} \vec v_{p-1} \\ - v_p \end{array}\right) = - \frac{\vec u_p}{|\vec u_p |}\label{cvuno} \ee defines the rotation \begin{equation}}\newcommand{\ee}{\end{equation} K_p ( \vec v_{p-1})\; \vec u_p = | \vec u_p | \; \; {\check e}_p \ee Note that we have changed the ($p-1$) parameters from $\vec u_p$ indicating its direction in terms of the ($p-1$) parameters in $\vec v_{p-1}$. With such a choice of $\vec v_{p-1}$ ($V_{p-1}$ not fixed yet) we can write \begin{equation}}\newcommand{\ee}{\end{equation} {\bar P}_{p+1} = K_{p+1} (| \vec u_p | \; {\check e}_p )\; H( H(V_{p-1},1)\; P_p\; H(V_{p-1},1)^t , 1) \ee where we have redefined $\; P_p\rightarrow K_p (\vec v_{p-1})^t\; P_p \; K_p(\vec v_{p-1})$. But according to \eqn{adjort} we can rewrite it as \begin{equation}}\newcommand{\ee}{\end{equation} {\bar P}_{p+1} = K_{p+1} (| \vec u_p | \; {\check e}_p )\; H({\bar P}_p , 1) \ee Inspection of this formula indicates an iterative process, the next step being to write the analogous expression for ${\bar P}_p$ and so on; after $p$ steps we get \begin{equation}}\newcommand{\ee}{\end{equation} {\bar P}_{p+1} = \prod_{l=1}^{\overleftarrow p}\; H\left( K_{l+1} ( |\vec u_l| {\check e}_l ), 1_{p-l} \right) \ee It is convenient to introduce the angular variables \begin{equation}}\newcommand{\ee}{\end{equation} |\vec u_l | = \sin \theta_l \;\; , \;\; 0 \leq\theta_l\leq \pi \ee and write the $SO(p+1)$ element in the final form \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} P_{p+1} &=& H( V_p , 1)^t\; {\bar P}_{p+1}\; H( V_p , 1)\cr {\bar P}_{p+1} &=& \prod_{l=1}^{\overleftarrow p}\; \exp\left( \theta_l\;A_{l,l+1}\right)\label{adjortfinal} \eea which displays explicitly the $p$ invariants $\{ \theta_l , l=1,\cdots,p \}$ under the adjoint action of $SO(p)$. Note that the number of parameters trivially matches: ${p\over 2} (p-1) + p = {p\over 2} (p+1)$, as should; this is the first of our results, to be used in the following. \subsection{Reduction of $SO(p,q)$ under $SO(p)\times SO(q)$.} Our starting point is again the right coset parametrization \cite{gil} \footnote{ An explicit derivation from the definition of pseudo unitary groups is given in Appendix A of \cite{lu1}. } \begin{equation}}\newcommand{\ee}{\end{equation} \Lambda_{p,q}(S,P_p,Q_q) = K_{p,q}(S)\; H(P_p,Q_q) \label{parortpq} \ee where $P_p (Q_q) \in (SO(p)( SO(q) )$ and the coset element is given by \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} K_{p,q}(S) &=& \exp\matriz{0}{N}{N^t}{0} = \matriz{(1 + S S^{t})^{\frac{1}{2}}}{S}{S^{t}}{(1 + S^{t}S)^{\frac{1}{2}}}\cr S &=& (N N^t )^{-{1\over2}} \sinh (N N^{t})^{1\over 2} N \;\; \in \Re^{p\times q} \label{cospseu} \eea Under an adjoint transformation \begin{equation}}\newcommand{\ee}{\end{equation} {}^h\Lambda_{p,q} = H(h_p,h_q )\;\Lambda_{p,q}\; H(h_p,h_q )^t \ee with $ h_p (h_q)\in SO(p)(SO(q))$, the parameters of $\Lambda_{p,q}$ transforms as \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} {}^h S &=& h_p \; S \; h_q{}^t \cr {}^hP_p &=& h_p \; P_p \; h_p^t\cr {}^hQ_q &=& h_q \; Q_q \; h_q^t \eea By following the strategy pursued in the past subsection we introduce two matrices $V_p, V_q$ belonging to $SO(p), SO(q)$ respectively and rewrite \eqn{parortpq} \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} \Lambda_{p,q} &=& \; H(V_p ,V_q )^t\;{\bar\Lambda}_{p,q} \;H(V_p ,V_q) \cr {\bar\Lambda}_{p,q} &=& exp\matriz{0}{V_p N V_q{}^t}{ (V_p N V_q{}^t )^t}{0} \; H(V_p P_p V_p{}^t , V_q Q_q V_q{}^t ) \eea As in \eqn{parV} we consider left coset parametrizations \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} V_p = H(V_{p-1}, 1) \;K_p(\vec v_{p-1})\cr V_q = H(V_{q-1},1)\;K_q(\vec v_{q-1}) \eea and try to totally fix $\;\vec v_{p-1}\;$ and $\;\vec v_{q-1}\;$ to put $N$ in a standard form. It turns out that it is possible to choose these vectors in such a way that \begin{equation}}\newcommand{\ee}{\end{equation} V_p \; N\; V_q{}^t \equiv H(V_{p-1} , 1)K_p(\vec v_{p-1})\; N K_q(\vec v_{q-1})^t\; H(V_{q-1}{}^t , 1) = \matriz{N_r}{0}{0}{n}\label{Nst} \ee where $\;N_r \in \Re^{(p-1)\times (q-1)}\;$ and $\;n\in \Re\;$. In fact if we write \begin{equation}}\newcommand{\ee}{\end{equation} N = \matriz{N'_r}{\vec n_{p-1}}{\vec n_{q-1}{}^t} {n'} \ee then is straightforward to verify that the gauge fixing condition $\vec n_{p-1}= \vec n_{q-1} = \vec 0$, i.e. \begin{equation}}\newcommand{\ee}{\end{equation} K_p(\vec v_{p-1})\;N K_q(\vec v_{q-1})^t = \matriz{N_r}{0}{0}{n} \ee holds if we choose \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} \vec v_{p-1} &\equiv& (1+ |\vec t_{p-1}|^2 )^{-\frac{1}{2}}\; \vec t_{p-1}\cr \vec v_{q-1} &\equiv& (1+ |\vec t_{q-1}|^2 )^{-\frac{1}{2}}\; \vec t_{q-1} \eea with $\;\vec t_{p-1} , \vec t_{q-1}\;$ satisfying \footnote{ This set of equations can be reduced to a system of two (quadratic) equations with two unknowns; the important thing for us is that solutions exist and define the change of variables \begin{eqnarray*} (\vec n_{p-1} ,\vec n_{q-1})\rightarrow (\vec v_{p-1} ,\vec v_{q-1}) \end{eqnarray*} } \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} \vec t_{p-1} &=& ( \vec n_{q-1}{}^t\vec t_{q-1} - n)^{-1}\;\left( \vec n_{p-1} - N'_r \vec t_{q-1}\right)\cr \vec t_{q-1} &=& (\vec n_{p-1}{}^t\vec t_{p-1} - n)^{-1}\left( \vec n_{q-1} - N'_r{}^t\vec t_{p-1}\right) \eea A final redefinition $\;N_r\rightarrow V_{p-1}{}^t\; N_r\; V_{q-1}\;$ leads to \eqn{Nst}. Finally (after reparametrizing $\; P_p \rightarrow K_p(\vec v_{p-1})^t\; P_p\; K_p(\vec v_{p-1})\; , \; Q_q \rightarrow K_q(\vec v_{q-1})^t\; Q_q\; K_q(\vec v_{q-1})\;$) we use \eqn{adjortfinal} to fix $V_{p-1}$ and $V_{q-1}$; the result can be recast in the form \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} \Lambda_{p,q} &=& \; H(V_p ,V_q )^t\;{\bar\Lambda}_{p,q} \;H(V_p ,V_q) \cr {\bar\Lambda}_{p,q} &=& \exp\left( \sum_{i=1}^{p-1}\sum_{j=1}^{q-1} N_{ij} \; S_{i,p+j} + n S_{p,p+q}\right)\; \prod_{k=1}^{\overleftarrow {p-1}}\; \exp(\theta_k A_{k,k+1})\; \prod_{k=1}^{\overleftarrow{q-1}}\;\exp(\bar\theta_k A_{p+k,p+k+1})\cr & &\label{adjpseu} \eea Again we verify the matching in the number of parameters $\; V_p, V_q, N_{ij}, n, \theta_k , \bar\theta_k\;$, \begin{equation}}\newcommand{\ee}{\end{equation} {p\over 2} (p-1)+ {q\over 2} (q-1) + (p-1)(q-1) + 1 + (p-1) + (q-1) = {1\over 2} (p+q) (p+q -1) \ee It is worth to say that the first term in \eqn{adjpseu} can be computed as in \eqn{cospseu} with $N$ as in the r.h.s. of \eqn{Nst}; however this is a formal expression for which, to our knowledge, only the ``minkowskian'' cases $p=1$ or $q=1$ admit an explicit form. \section{The unitary groups} \setcounter{equation}{0} The treatment of these groups parallels that made in the case of the orthogonal ones, with some additional complications due to the complex character of them. As before we start considering \subsection{Reduction of $SU(p+1)$ under $U(p)$.} An element of $SU(p+1)$ can be written as \begin{equation}}\newcommand{\ee}{\end{equation} P_{p+1}(\vec u_p , U_p ) = K_{p+1}(\vec u_p )\; H(U_p, u_p^* ) = K_{p+1}(\vec u_p )\; H(P_p,1)\; \exp\left( i\;\phi^p T_p\right)\label{cosun} \ee where $\; u_p\equiv\det U_p=\exp(ip\phi_p )\;,\;\vec u_p\in\mbox{{\bf C}}^p\; ,\; U_p\in U(p),P_p\in SU(p)\;$, and we have introduced a convenient basis $\;\{ T_k = \sum_{l=1}^k ( E_{ll} - E_{k+1,k+1} ),\; k=1,\cdots,p\}\;$ in the Cartan subalgebra of $su(p+1)$. The right coset element in \eqn{cosun} belonging to $SU(p+1)/U(p)\sim CP^p$ is given by \cite{gil} \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} K_{p+1}(\vec u_p ) &=& \matriz{1 - (1 + u_{2p+1} )^{-1}\; \vec u_p\vec u_p{}^\dagger} {\vec u_p}{-\vec u_p{}^\dagger}{u_{2p+1}} \cr 1 &=& \vec u_p{}^\dagger \vec u_p + (u_{2p+1})^2 \eea The adjoint action under $\; H_p \in U(p)\;$ is ($\;h_p \equiv \det H_p$) \begin{equation}}\newcommand{\ee}{\end{equation} {}^H P_{p+1} = H( H_p, h_p^*)\; P_{p+1}\; H(H_p ,h_p^*)^\dagger \;\longleftrightarrow\; \left\{ \begin{array}{l} {}^V\vec u_p = h_p\; H_p \; \vec u_p \\ {}^H P_p = H_p\; P_p \; H_p^\dagger \\\; {}^H \phi^p = \phi^p \end{array}\right. \label{adjtrpsu} \ee As before we pick an arbitrary $V_p \in U(p)$ left coset parametrized \begin{equation}}\newcommand{\ee}{\end{equation} V_p = H( V_{p-1} , v_{p-1}^*)\; K_p (\vec v_{p-1})\; \exp( i\beta_p)\;\;\; ,\;\; V_{p-1}\in U(p-1) \ee and write \eqn{cosun} with the help of \eqn{adjtrpsu} as \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} P_{p+1} &=& H(V_p , v_p^*)^\dagger\; {\bar P}_{p+1} \; H(V_p,v_p{}^*)\cr {\bar P}_{p+1} &=& K_{p+1}(v_p V_p\vec u_p )\; H(V_p P_p V_p{}^\dagger ,1)\; \exp(i\;\phi^p T_p ) \label{adjuni} \eea By choosing $\vec v_{p-1}$ and ${\beta_p}$ ($V_{p-1}$ free) as \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} \left( \begin{array}{c} \vec v_{p-1} \\ - v_{2p-1} \end{array}\right) &=& - \frac{(\vec u_p^*)^p}{|(\vec u_p)^p |}\; \frac{\vec u_p}{|\vec u_p |} \cr \exp (i\beta_p ) &=& \left( \frac{(\vec u_p^* )^p}{|(\vec u_p{} )^p |} \right)^\frac{1}{p+1}\label{cvdos} \eea we have \begin{equation}}\newcommand{\ee}{\end{equation} v_p \; V_p \; \vec u_p = |\vec u_p| \; {\check e}_p \ee and identifying $V_{p-1}$ as the $SU(p-1)$ matrix corresponding to the $P_p$ decomposition in \eqn{adjuni} ( previous redefinition $\; P_p\rightarrow K_p\; (\vec v_{p-1})^\dagger P_p \; K_p(\vec v_{p-1})\;$ ) we get \begin{equation}}\newcommand{\ee}{\end{equation} {\bar P}_{p+1} = K_{p+1}(|\vec u_p | {\check e}_p )\; H({\bar P}_p ,1)\;\exp(i\;\phi^p T_p ) \ee By repeating the analysis for ${\bar P}_p $ and after $p$ steps we arrive to the final result \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} P_{p+1} &=& H( V_p , v_p^*)^\dagger\; {\bar P}_{p+1}\; H( V_p ,v_p^* )\cr {\bar P}_{p+1} &=& \prod_{l=1}^{\overleftarrow p}\; \exp\left( \theta_l\; A_{l,l+1}\right)\; C_{p+1}(\Phi ) \label{adjunifinal} \eea It differs from \eqn{adjortfinal} from the unitary character of $V_p$ and the comparison of the arbitrary Cartan element $\;C_{p+1}(\Phi )=\exp(i \sum_{l=1}^p \;\phi^l T_l )\;$ at right in ${\bar P}_{p+1}$. \subsection{Reduction of $SU(p,q)$ under $S(U(p)\times U(q))$.} In order not to be repetitive we will skip some steps in what follows. An arbitrary element in $SU(p,q)$ can be left coset decomposed under $\;S(U(p)\times U(q))\;$ as \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} \Lambda_{p,q}(S,U_p,U_q) &=& K_{p,q}(S)\; H(U_p,U_q)\;\;\; ,\;\; u_p = u_q{}^* \cr K_{p,q}(S) &=& \matriz{(1 + S S^\dagger)^{\frac{1}{2}}}{S}{S^\dagger}{(1 + S^\dagger S)^{ \frac{1}{2}} } = \exp \matriz{0}{N}{N^\dagger}{0}\cr S &=& (N N^\dagger )^{-{1\over2}} \sinh (N N^\dagger)^{1\over 2} N \;\;\; ,\;\; S, N\in {\cal C}^{p\times q}\label{parunipq} \eea The adjoint action under $\; H(h_p,h_q ) \in S(U(p)\times U(q) )\;$ is \begin{equation}}\newcommand{\ee}{\end{equation} {}^H \Lambda_{p,q} = H(h_p,h_q )\;\Lambda_{p,q} H(h_p ,h_q )^\dagger \longleftrightarrow \left\{\begin{array}{l} \;{}^H S = h_p \; S \; h_q{}^\dagger \cr {}^H U_p = h_p \; U_p \; h_p^\dagger\cr {}^H U_q = h_q \; U_q \; h_q{}^\dagger \end{array}\right. \ee Two matrices $V_p, V_q$ belonging to $U(p), U(q)$ respectively with $v_p v_q =1$ are introduced and following similar steps as in Subsection $2.2$ we find that it is possible to fix $N$ in the way \begin{equation}}\newcommand{\ee}{\end{equation} N = \matriz{N_r}{0}{0}{n} \ee where now $N_r \in {\cal C}^{(p-1)\times(q-1)}$ and $n\in \Re $. Then by using the results of the past subsection we get the final result for the parametrization \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} \Lambda_{p,q} &=& \; H(V_p, V_q)^\dagger \;{\bar \Lambda}_{p,q} \;H(V_p ,V_q) \cr {\bar \Lambda}_{p,q} &=& \exp \matriz{0}{N}{N^\dagger}{0}\; \overleftarrow{\prod_{l=1}^{p-1}}\;\exp\left(\theta_l\; A_{l,l+1}\right)\; \overleftarrow{\prod_{l=1}^{q-1}}\;\exp\left({\bar\theta}_l\; A_{p+l,p+l+1}\right)\;C(\Phi) \label{adjpsunifin} \eea where $\; H(V_p, V_q)\in S(U(p)\times U(q))\;$ and we denote by $\;C(\Phi)\;$ an arbitrary element in the Cartan subalgebra of the Lie algebra of $S(U(p)\times U(q))$. \subsection{Decomposition under the maximal torus} Some times is useful to have the adjoint decomposition of unitary groups under the Cartan subalgebra. We will work out for definiteness the case of $SU(n+1)$; the non compact versions differ as usual by signs in the coset elements and Wick rotations of some compact generators. To this end we begin by searching for the coset decomposition of $SU(n+1)$ under $C(SU(n+1))$; from \eqn{cosun} \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} P_{n+1}(\vec u_n , U_n ) &=& K_{n+1}(\vec u_n )\; H(U_n, u_n^* )\cr K_{n+1}(\vec u_n ) &=& \exp\theta_n\matriz{0}{\check{r}_n}{-\check{r}_n{}^\dagger}{0} = \matriz{1 - (1 + u_{2n+1} )^{-1}\; \vec u_n\vec u_n{}^\dagger}{\vec u_n}{-\vec u_n{}^\dagger}{u_{2n+1}}\cr |\vec u_n|^2 + (u_{2n+1})^2 &=& 1 \;\;\;, \;\;\vec u_n = \sin\theta_n \;{\check r}_n\;\;\; , \;\;{\check r}_n{}^\dagger {\check r}_n =1\;\;\; ,\;\; \theta_n \in [0,\pi]\label{ec} \eea By introducing \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} U_n &=& P_n\;\matriz{1_{n-1}}{0}{0}{u_n} \;\; , \;\; P_n \in SU(n)\cr u_n &\equiv& \det U_n = \exp(i\varphi^n) \eea we can rewrite \eqn{ec} as \begin{equation}}\newcommand{\ee}{\end{equation} P_{n+1}(\vec u_n ,U_n ) = K_{n+1}(\vec u_n )\; H(P_n ,1)\;\exp(i\;\varphi^n H_n) \ee where this time is convenient to introduce the basis $\;\{ H_k = E_{kk} - E_{k+1,k+1} ,k=1,\cdots, n\}\;$ in the Cartan subalgebra of $su(n+1)$. By repeating with $P_n$ and iterating we get \begin{equation}}\newcommand{\ee}{\end{equation} P_{n+1} = \overleftarrow{\prod_{l=1}^n} \; \matriz{K_{l+1}(\vec u_l)}{0}{0}{1_{n-l}} \;\exp(i\;\vec\varphi \cdot {\vec H}) \ee Now let us pick an element $C_{n+1}(\vec\alpha )\equiv \exp (\alpha^i H_i)\in C(SU(n+1))$ and write as usual \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} P_{n+1} &=& C_{n+1}(\vec\alpha )^\dagger \;\overleftarrow{\prod_{l=1}^{n}} \left( C_{n+1}(\vec\alpha )\;\matriz{K_{l+1}(\vec u_l)}{0}{0}{1_{n-l}}\; C_{n+1}(\vec\alpha )^\dagger\right)\; C_{n+1}(\vec\varphi)\;C_{n+1}(\vec\alpha )\cr & & \label{unicarini} \eea We see that the $(\varphi^l )$ variables are invariant; we must choose the ${\alpha^i}$'s to kill parameters in the productory. It is easy to show that the versors $\check{r}_l$ get transformed as \begin{equation}}\newcommand{\ee}{\end{equation} {}^\alpha\check{r}_l = \exp (-i\tilde\alpha_{l+1} )\; \exp( i\sum_{i=1}^{l}\tilde\alpha_i E_{ii})\;\check{r}_n\;\;\; ,\;\; l=1,\cdots,n \ee where \begin{equation}}\newcommand{\ee}{\end{equation} {\tilde\alpha}_k = \left\{ \begin{array}{ll} \alpha_1 & \;\; if\;\; k=1\cr -\alpha_{k-1} + \alpha_k & \;\; if\;\; k=2,\ldots,n \cr -\alpha_n & \;\; if\;\; k=n+1 \end{array} \right. \ee Then we can put the phases of the $({}^\alpha\check{r}_l )^l$ components, $l=1,\cdots,n$, to zero by choosing the $\alpha's$ such that \footnote{Equations \eqn{kcm} can be formally solved by $\;\; \alpha^i = - {K^{-1}}^{i}{}_j \; phase(\check{r}_j )^j$ where $K$ is the Killing-Cartan matrix of the $A_n$ algebra; it is probable that this fact does not be an accident but occurs in other cases $G/C(G)$. } \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} 2 \alpha_1 -\alpha_2 &=& - phase(\check{r}_1 )^1\cr - \alpha_1 + 2\alpha_2 -\alpha_3 &=& -phase(\check{r}_2)^2\cr \vdots & &\vdots\cr -\alpha_{n-1}+2\alpha_n &=& -phase(\check{r}_n )^n \label{kcm} \eea In other words, from \eqn{unicarini} we get the final result \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} P_{n+1} &=& C_{n+1}(\vec\alpha)^\dagger\; {\bar P}_{n+1}\; C_{n+1}(\vec\alpha)\cr {\bar P}_{n+1} &=& \overleftarrow{\prod_{l=1}^{n}} \matriz{K_{l+1}(\vec u_l)}{0}{0}{1_{n-l}}\;C_{n+1}(\vec\varphi )\label{unicarfin} \eea with the constraints implied by \eqn{kcm}: $\;(\check{r}_l)^l \in \Re \; ,\; l=1,\ldots, n$. \section{The decomposition of $Sl(n)$ under $SU(n)$.} \setcounter{equation}{0} This is one of the two irreducible riemannian cases \footnote{ The other one is $SU^*(2n)/USp(2n)$ and will not be considered here.} in the sense that the coset element is not of the form $\exp\matriz{0}{N}{\pm N^\dagger}{0}\;$ for some matrix $N$ (off-diagonal cosets). We start from the well-known coset decomposition under $SO(n)$ of any unimodular real $n\times n$ matrix \begin{equation}}\newcommand{\ee}{\end{equation} g_n = S_n \; P_n\;\; ,\; S_n{}^t = S_n \;\;\; ,\;\; P_n \in SO(n) \ee Also $S_n$ is positive definite and $\det S_n = 1$. But we know from elementary linear algebra that any such a matrix is diagonalizable by an orthogonal one $Q_n$ completely determined \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} S_n &=& Q_n{}^t \; Diag(\lambda_1{}^2 , \ldots, \lambda_n{}^2 ) \; Q_n\cr \prod_{i=1}^{n}\lambda_i &=& 1\;\;\; ,\;\; \lambda_1\geq \lambda_2\geq \dots \geq\lambda_n\geq 0 \eea from where after a redefinition $P_n \rightarrow Q_n{}^t P_n Q_n$ we get \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} g_n &=& Q_n{}^t \;{\bar g}_n \; Q_n \;\;\; , Q_n\in SO(n)\cr {\bar g}_n &=& Diag(\lambda_1{}^2 , \ldots, \lambda_n{}^2 )\; P_n \label{slnort} \eea Analogous steps using well known results yield the complexification of \eqn{slnort}, namely the decomposition of $Sl(n, \cal C)$ under $SU(n,\cal C)\;$ that we quote without proof \begin{eqnarray}} \newcommand{\eea}{\end{eqnarray} g_n &=& V_n{}^\dagger \;{\bar g}_n \; V_n \;\;\; , \;\; V_n \in SU(n, \cal C )\cr {\bar g}_n &=& C_{n}(\vec\alpha) \; \overleftarrow{\prod_{l=1}^{n-1}} \matriz{K_{l+1}(\vec u_l )}{0}{0}{1_{n-1-l}}\; C_{n}(\vec\beta)\label{slnuni} \eea where $\; C_n \in C(SU(n,{\cal C} ))\;$ and the vectors $\vec u_l$ are constrained by $\; (\check{r}_l)^l \in \Re\; $ as in Section 3.3. \section{Conclusions} We have obtained in this paper adjoint parametrizations defined by \eqn{adj} w.r.t. maximally compact subgroups for a large set of non-compact groups, the classically riemannian cosets. The constructive procedure used allows to extend them straightforwardly to non riemannian decompositions, for example $\;SO(p+n,q)\;$ under $\;SO(n,q)\;$. Few words about symplectic groups: these groups can be treated in the same way as made here; the fact that $\; USp(2p, 2q) \sim U(2p, 2q; {\cal C} )\bigcap Sp(2p+2q;{\cal C})\;$ seems to suggest that the corresponding decomposition under $\;USp(2p)\times USp(2q)\;$ could follow from replacing in \eqn{adjpsunifin} $E$ by $Z$ generators \footnote{ See reference \cite{gil}, chapter 5 for definitions. } and respective Cartan subalgebras, but we have not checked this. We remark the locality of the parametrizations obtained; the changes of variables needed to carry out the job are singular in some points of the group manifold as can be seen by direct inspection of \eqn{cvuno} , \eqn{cvdos} for example. Possible applications in physical problems of these parametrizations are in the context of GWZW models as models of strings moving on background fields. Equation \eqn{adjpseu} can be used to treat systematically all the models in \cite{bs1} where the lowest dimensional cases were considered. Also the measure on the group can be computed straightforwardly through the Maurer-Cartan forms without introducing Fadeed-Popov ghost due to constraints because they were solved once for all. For $p+q= 2$ ($A_1$ algebras) parametrization \eqn{adjpsunifin} is widely known; for $p=2, q=1$ was introduced in \cite{lu1}; it allows to extend the study of coset models in the search of physically relevant string backgrounds represented by exact conformal field theories. \section*{Acknowledgements} It is a pleasure to thank to Loriano Bonora for useful correspondence.
\section{Introduction and statements of results} The concept of Murasugi sum (for the definition, see Section 2) of Seifert surfaces in the $3$-sphere $S^{3}$ was introduced by K. Murasugi, and it has been playing important roles in the studies of Seifert surfaces and links. The Murasugi sum is known to be natural in many senses, and in particular the following is known. (We say that a Seifert surface $R$ is a {\it fiber surface} if $\partial R$ is a fibered link and $R$ realizes the fiber.) \begin{thm}[\mbox{\cite[Theorem 3.1]{G}}] \label{thm:fiber-sum} Let $R$ be a Murasugi sum of $R_{1}$ and $R_{2}$. Then $R$ is a fiber surface if and only if both $R_{1}$ and $R_{2}$ are fiber surfaces. \end{thm} On the other hand, the concept of alternating link has also been important in knot theory. It has been known that there are some relationships between alternating diagrams and the Seifert surfaces obtained by applying Seifert's algorithm to them. For example, if a link diagram $D$ is alternating, then the Seifert surface obtained from $D$ by the algorithm is of minimal genus, \cite{C, Mu1958}. In \cite{G}, D. Gabai gave a geometric proof to the following theorem, which also follows from \cite{Mu} and \cite{St}. Note that if $L$ is fibered, then minimal genus Seifert surfaces for $L$ are unique up to isotopy and the fiber is realized by the minimal genus surface. \begin{thm}[\mbox{\cite[Theorem 5.1]{G}}] \label{thm:fiber-hopf} Let $L$ be an oriented link with an alternating diagram $D$. $L$ is a fibered link if and only if the surface $R$ obtained by applying Seifert's algorithm to $D$ is connected and (obviously) desums into a union of Hopf bands. \end{thm} We say that a Seifert surface $R (\subset S^3)$ {\it desums} into $R_{1},\ldots , R_{n}$ if $R$ is a Murasugi sum of them. Especially, if $R$ is obtained by successively plumbing (i.e., 4-Murasugi summing) finite number of Hopf bands to a disk, we call $R$ a {\it Hopf plumbing}. Actually, the \lq only if' part of Theorem \ref{thm:fiber-hopf} can be strengthened as in the following theorem, which follows from Propositions \ref{prop:para-deplumb} and \ref{prop:alt-para}. \begin{thm} \label{thm:alt-fiber} Let $L$ be an oriented link with an alternating diagram $D$. $L$ is a fibered link if and only if the surface $R$ obtained by applying Seifert's algorithm to $D$ is a Hopf plumbing. Moreover, $R$ is a fiber surface if and only if $R$ is deformed into a disk by successively cutting one of a pair of \lq parallel bands' (defined in Section 5). \end{thm} In \cite{A}, C. Adams et al. generalized the concept of alternating links and introduced the concept of almost alternating links. A diagram $D$ in $S^2$ is called {\it almost alternating} (resp. {\it $2$-almost alternating}) if $D$ becomes an alternating diagram after one crossing change (resp. two crossing changes). A link $L$ in $S^3$ is called {\it almost alternating} if $L$ is not alternating but admits an almost alternating diagram. If $D$ is an almost alternating diagram, the specific crossing to change is called the {\it dealternator} and we call the other crossings the {\it alternators}. In this paper, we extend Theorems \ref{thm:fiber-hopf} and \ref{thm:alt-fiber} to almost alternating links. Note that almost alternating diagrams, however, do not always yield a minimal genus Seifert surface via Seifert's algorithm. Our first result is as follows: \begin{thm} \label{thm:main} Let $D$ be an almost alternating diagram, and $R$ a Seifert surface obtained by applying Seifert's algorithm to $D$. Then, $R$ is a fiber surface if and only if $R$ is connected and desums into a union of Hopf bands. \end{thm} In section 5, we show a stronger version of Theorem \ref{thm:main} as below, by using Corollary \ref{cor:algorithm} obtained from the arguments in the proof of Theorem \ref{thm:main}. \begin{thm} \label{thm:hopf-plumbing} Let $R$ be a Seifert surface obtained by applying Seifert's algorithm to an almost alternating diagram. Then, $R$ is a fiber surface if and only if $R$ is a Hopf plumbing. \end{thm} As a corollary of the proof of Theorem \ref{thm:main}, we obtain a practical algorithm to determine whether or not a given almost alternating diagram yields a fiber surface via Seifert's algorithm. We use this to prove Theorem 1.5. We say that a diagram $D$ is {\it unnested} if $D$ has no Seifert circle which contains another circle in both of its complementary region. Otherwise we say $D$ is {\it nested}. \begin{cor}\label{cor:algorithm} Let $D$ be an almost alternating diagram and $R$ a Seifert surface obtained from $D$ by Seifert's algorithm. Then $R$ is a fiber surface if and only if $R$ is connected and desums into a union of Hopf bands by repeating of the following decompositions;\\ (1) a Murasugi decomposition along a nested Seifert circle,\\ (2) a prime decomposition, and\\ (3) Murasugi decompositions of type (A) and (B) in Figure 1.1, where each decomposition yields Seifert surfaces with first Betti numbers smaller than that of $R$. \end{cor} \figbox{Figure 1.1} \noindent {\it Proof. } In the proof of Theorems \ref{thm:fiber-hopf} (see \cite[p.533]{G}) and \ref{thm:main}, we explicitly show how we can desum such $R$ into surfaces of smaller first Betti numbers. All necessary decompositions are covered in the above three. \fpf In \cite{Ha}, J. Harer proved that every fiber surface in $S^3$ results from a disk by a sequence of elementary changes as follows: (a) plumb on a Hopf band, \par (b) deplumb a Hopf band, and\par (c) perform a Dehn twist about a suitable unknotted curve in the fiber.\par Then he asked whether changes of either type (b) or (c) can be omitted, and any fiber surface can be realized only using changes of the remaining two types. So it is worthy presenting the following partial affirmative answer as a corollary, which immediately follows from Theorem \ref{thm:hopf-plumbing} and Propositions 5.1 and 5.2. \begin{cor}\label{cor:only-plumb} Let $R_1$ and $R_2$ be any fiber surface obtained by applying Seifert's algorithm to an alternating or almost alternating diagram. Then $R_1$ and $R_2$ can be changed into each other by plumbing and deplumbing Hopf bands. \end{cor} We say that a Hopf hand $B$ is {\it positive} (resp. {\it negative}) if the linking number of $\partial B$ is $1$ (resp. $-1$). By the following fact together with an observation of the way fiber surfaces deplumb in the proof of Theorem \ref{thm:main} (see Section 4) and Theorem \ref{thm:hopf-plumbing} (see Section 5), we have the following Corollary: \begin{cor}\label{cor:posi-hopf} Let $D$ be an unnested almost alternating diagram such that the sign of the dealternator is negative. Suppose the surface $R$ obtained from $D$ by Seifert's algorithm is a fiber surface. Then $R$ is a plumbing of positive Hopf bands. \end{cor} \begin{fact} Suppose a diagram $D$ is unnested. Then $D$ is alternating (resp. almost alternating) if and only if all the crossings of $D$ have the same sign (resp. the same sign except exactly one crossing). \end{fact} This paper is organized as follows; Section 2 is for preliminaries. In Section 3, we give an example for our theorem. We also show that our theorem can not be extended to $2$-almost alternating diagrams, i.e., (1) we recall Gabai's example (in \cite{G}) of a $2$-almost alternating diagram for a link whose Seifert surface obtained by Seifert's algorithm is a fiber surface that is not a nontrivial Murasugi sum, and (2) we give examples of $2$-almost alternating diagrams for knots whose Seifert surfaces obtained by Seifert's algorithm are fiber surfaces that are not Hopf plumbing. In Sections 4 and 5, we prove Theorems~\ref{thm:main} and ~\ref{thm:hopf-plumbing} respectively. \section{Preliminaries} For the definitions of standard terms of sutured manifolds, see \cite[p.520]{G}. We say that a sutured manifold $(M,\gamma)$ is a {\it product sutured manifold} if $(M,\gamma)$ is homeomorphic to $(R \times I, \partial R \times I)$ with $R_{+}(\gamma)=R \times \{ 1\}, R_{-}(\gamma)=R \times \{ 0\}$, where $R$ is a compact oriented surface with no closed components and $I$ is the unit interval $[0,1]$. The {\it exterior} $E(L)$ of a link $L$ in $S^{3}$ is the closure of $S^{3}-N(L;S^{3})$. If $R$ is a Seifert surface for $L$, we may assume $R\cap E(L)$ is homeomorphic to $R$, and often abbreviate $R\cap E(L)$ as $R$. Let $R$ be a Seifert surface for $L$ in $S^3$. The product sutured manifold $(M,\gamma)=(R \times I, \partial R \times I)$ is called the sutured manifold {\it obtained} from $R$ and the sutured manifold $(N,\delta)=(E(L) - {\rm Int}M, \partial E(L) - {\rm Int}\gamma)$ is the {\it complementary} sutured manifold for $R$ (or for $(M, \gamma)$). Note that $R$ is a fiber surface if and only if the complementary sutured manifold for $R$ is a product sutured manifold. A {\it product decomposition} \cite{G} is a sutured manifold decomposition $$(M_1,\gamma_1) \overset{B}{\longrightarrow} (M_2,\gamma_2),$$ where $B$ is a disk properly embedded in $M_1$ such that $B \cap s(\gamma_1)=$ ($2$ points), $M_2 = M_1 - {\rm Int}N(B)$ and that $s(\gamma_2)$ is obtained by extending $s(\gamma_1) - {\rm Int}N(B)$ in the natural way (Figure 2.1 (a)). The disk $B$ is called a {\it product disk}. Dually, {\it $C$-product decomposition} is the operation $$(M_1, \gamma_1) \overset{E}{\longrightarrow} (M_2, \gamma_2),$$ where $E$ is a disk properly embedded in $S^3 - {\rm Int}M_1$ such that $E \cap s(\gamma_1)=$ ($2$ points), $M_2$ is obtained from $M_1$ by attaching the 2-handle $N(E)$ and that $s(\gamma_2)$ is obtained by extending $s(\gamma_1) - {\rm Int}N(E)$ in the natural way (Figure 2.1 (b)). The disk $E$ is called a {\it $C$-product disk}. \figbox{Figure 2.1} \noindent {\bf Definition.} Let $R$ be a Seifert surface for a link $L$. We say that {\it $R$ has a product decomposition} if there exists a sequence of $C$-product decompositions $$(R \times I, \partial R \times I)=(M_0, \gamma_0) \overset{E_1}{\longrightarrow} (M_1, \gamma_1) \overset{E_2}{\longrightarrow} \cdots \overset{E_p}{\longrightarrow} (M_p, \gamma_p),$$ where the complementary sutured manifold for $(M_p, \gamma_p)$ is a union of $3$-balls each with a single suture. As a criterion to detect a fiber surface, Gabai has shown the following: \begin{thm}[\mbox{\cite[Theorem 1.9]{G}}] \label{thm:fiber-pd} Let $L$ be an oriented link in $S^{3}$, and $R$ a Seifert surface for $L$. Then, $L$ is a fibered link with fiber $R$ if and only if $R$ has a product decomposition. \end{thm} We note that in Section 4, the existence of a $C$-product decomposition $(M_0, \gamma_0) \overset{E_1}{\longrightarrow} (M_1, \gamma_1)$ together with the $C$-product disk $E_1$ is important. \medskip \noindent {\bf Definition.} A surface $R\,(\subset S^{3}$) is a {\it $2n$-Murasugi sum} of two surfaces $R_{1}$ and $R_{2}$ in $S^{3}$ if the following conditions are satisfied; \begin{enumerate} \item $R=R_{1}\underset{\Delta}{\cup}R_{2},$ where $\Delta$ is a $2n$-gon, i.e., $\partial \Delta=\mu_{1}\cup\nu_{1}\cup\ldots\cup\mu_{n}\cup\nu_{n}$ (possibly $n=1$), where $\mu_{i} ($resp. $\nu_{i}$) is an arc properly embedded in $R_{1} ($resp. $R_{2}$). \item There exist $3$-balls $B_{1}$ and $B_{2}$ in $S^{3}$ such that:\\ (i) $B_{1}\cup B_{2}=S^{3},\,B_{1}\cap B_{2}=\partial B_{1}=\partial B_{2}=S^{2}$ : a $2$-sphere,\\ (ii)$R_{1}\subset B_{1}, R_{2}\subset B_{2}$ and $R_{1}\cap S^{2}=R_{2}\cap S^{2}=\Delta.$ \end{enumerate} The $2$-Murasugi sum is known as the connected sum, and the $4$-Murasugi sum is known as the plumbing. \figbox{Figure 2.2} Concerning alternating and almost alternating tangles, we can confirm the following facts. \begin{fact}\label{fact:tangle-sum} Suppose a link diagram $D$ is a tangle sum of two tangle diagrams $D_1$ and $D_2$. If $D$ is alternating, then both $D_1$ and $D_2$ are alternating. And if $D$ is almost alternating, then one of them, say, $D_1$ is alternating and $D_2$ is almost alternating. \end{fact} \begin{fact}\label{fact:tangle} By connecting neighboring strands running out of an alternating (resp. almost alternating ) tangle diagram, we obtain an alternating (resp. almost alternating ) link diagram. See Figure 2.3. \end{fact} \figbox{Figure 2.3} Then by these two facts, we can confirm the following propositions. Let $R$ be a Seifert surface obtained by applying Seifert's algorithm to a diagram $D$. \begin{prop}\label{prop:prime} If an almost alternating diagram $D$ is a connected sum of two diagrams, then one of them, say, $D_1$ is alternating and the other, say, $D_2$ is almost alternating. The Seifert surface $R$ is a $2$-Murasugi sum of $R_1$ and $R_2$, where $R_i$ is obtained from $D_i$. \end{prop} \begin{prop}\label{prop:nest} Suppose that an almost alternating diagram $D$ has a nested Seifert circle $C$. Then, along the disk bounded by $C$, $R$ is a Murasugi sum of $R_1$ and $R_2$, where $R_1$ (resp. $R_2$) is obtained from an alternating (resp. almost alternating) diagram. \end{prop} \begin{prop}\label{prop:induction} Suppose that $R$ desums into two surfaces $R_1$ and $R_2$ as illustrated in Figure 1.1, where the left figures in (A) and (B) are both almost alternating. Then $R_i$ $(i=1,2)$ is obtained from an alternating or almost alternating diagram. \end{prop} \section{Examples} In this section, we present some examples. Example 3.1 is for Theorems \ref{thm:main} and \ref{thm:hopf-plumbing}. Examples 3.2 and 3.3 show our Theorems \ref{thm:main} and \ref{thm:hopf-plumbing} can not be extended to $2$-almost alternating diagrams. For the names of knots, refer to Rolfsen's book \cite{R}. \medskip \noindent {\bf Example 3.1.} Figure 3.1 depicts an almost alternating diagram for the knot $10_{151}$, together with a fiber surface $R$ obtained by Seifert's algorithm. We can observe that $R$ desums into a union of Hopf bands and is a Hopf plumbing. \figbox{Figure 3.1} \medskip \noindent {\bf Example 3.2.} Let $R$ be the Seifert surface obtained by applying Seifert's algorithm to the oriented pretzel link diagram of type $(2,-2,2p)$ as in Figure 3.2, where $p\neq 0$. $R$ is a fiber surface but does not desum into a union of Hopf bands. \medskip We note that this example has been known in \cite{G} as a fiber surface for a link which does not admit a non-trivial Murasugi sum. \figbox{Figure 3.2} \noindent {\bf Example 3.3.} Figure 3.3 depicts 2-almost alternating diagrams for the knots $9_{42}$, $9_{44}$ and $9_{45}$. By applying Seifert's algorithm to them, we obtain fiber surfaces, which are not Hopf plumbings. This can be shown by the following proposition and direct calculations of genera and the Conway polynomials of these knots. \setcounter{thm}{3} \begin{prop}[\mbox{\cite[Theorem 3]{MM}}] If a fibered knot $K$ of genus $2$ can be constructed by plumbing Hopf bands, then the Conway polynomial $\nabla_K(z)$ of $K$ satisfies the following; $$ \nabla_K(z) \neq \begin{cases} 1 + c_{1} z^2 + z^4 & for \ c_1 = 0 \ {\rm mod} \ 4,\cr 1 + c_{1} z^2 - z^4 & for \ c_1 = 2 \ {\rm mod} \ 4. \cr \end{cases} $$ \end{prop} \figbox{Figure 3.3} \section{Proof of Theorem 1.4} Since the \lq if\rq \ part is shown by Theorem \ref{thm:fiber-sum}, we show the \lq only if\rq \ part. Let $D$ be an almost alternating diagram for a link $L\,(\subset S^3)$ on the {\it level $2$-sphere} $S^2$ and let $R$ be a Seifert surface obtained by applying Seifert's algorithm to $D$. Note that if a diagram $D$ is unnested, then Seifert's algorithm uniquely yields a Seifert surface. We say that a Seifert surface $R$ is {\it flat} if $R$ is obtained from an unnested diagram and thus lies in $S^2$ except in the neighborhood of each crossing. Suppose that $R$ is a fiber surface. Since any fiber surface is connected, we can assume $D$ is connected. Suppose $D$ is nested. Then, by Theorem \ref{thm:fiber-sum}, $R$ desums into fiber surfaces $R_1$ and $R_2$. Moreover, by Proposition \ref{prop:nest}, one of them, say, $R_1$ is obtained from an alternating diagram and $R_2$ from an almost alternating diagram. By Theorem \ref{thm:fiber-hopf}, $R_1$ desums into a union of Hopf bands. Therefore, we may assume that $D$ is unnested. Similarly, by Proposition \ref{prop:prime}, we may assume that $D$ is prime, and in particular, reduced. Now we prove the theorem by induction on the first Betti number $\beta_{1}$ of $R$, where $R$ is a fiber surface obtained by applying Seifert's algorithm to a connected unnested prime almost alternating diagram $D$. If $\beta_{1}=1$, then $R$ is an unknotted annulus and $D$ has $n$ crossings which are of the same sign except exactly one crossing. Note that $R$ is a fiber surface if and only if $n=4$, in which case $R$ is a Hopf band. Hence we have the conclusion. Then we assume that the theorem holds when $\beta_1(R) < k$ and prove the theorem for $R$ with $1\le\beta_1(R) = k$. The main method of the proof is to examine the $C$-product disk for the sutured manifold obtained from $R$ and grasp a local picture where we can desum $R$ into surfaces $R_1$ and $R_2$ obtained by the algorithm with smaller first Betti numbers. In each case, it is easy to confirm that $D_{i}\,(i=1,2)$ is an alternating or almost alternating diagram, that is, they satisfy the assumption of the induction (see Corollary \ref{cor:algorithm} and Proposition \ref{prop:induction}). Let $(M, \gamma)$ be the sutured manifold obtained from $R$. We identify $s(\gamma)$ as $L$. Let $E$ be a $C$-product disk for $(M, \gamma)$, i.e., $E$ is properly embedded in $S^3 - {\rm Int}M$ so that $E \cap L =$ ($2$ points). We may suppose that $E$ is non-boundary-parallel, and assume that $|E\cap S^2|$ is minimal among all such disks. Further, we may assume by isotopy that $\partial E \cap L$ occurs only in small neighborhoods of the crossings of $D$. Similarly, we can assume that $\partial E\cap S^{2}$ occurs only in small neighborhoods of the crossings. For convenience, we say that {\it $\partial E\cap L$ and $\partial E\cap S^{2}$ occur at the crossings.} {\bf Case A.} $E\cap S^{2}=\emptyset.$ If $\partial E\cap L$ occurs at one crossing, then $E$ is boundary parallel, a contradiction. Thus, we suppose that $\partial E\cap L$ occurs at two crossings (see Figure 4.1). If both crossings are alternators, we see that $R$ is a plumbing of flat surfaces, one of which is obtained from an unnested almost alternating diagram and has first Betti number smaller than $k$. If one crossing is the dealternator, we also see that $R$ is a plumbing of surfaces, one of which is compressible and hence not a fiber surface, a contradiction to Theorem \ref{thm:fiber-sum}. \figbox{Figure 4.1} {\bf Case B.} $E\cap S^{2}\neq\emptyset.$ Label the crossings with $\fr0, \fr1, \ldots , \fr{w-1}$ so that the dealternator has $\fr0$. By standard innermost circle argument, we may assume, by the minimality of $|E \cap S^2|$, that $E\cap S^{2}$ consists of arcs. Let $\alpha$ be an arc of $E\cap S^{2}$. By assumption, each endpoint of $\alpha$ lies in a neighborhood of a crossing and hence is accordingly labeled. Then the {\it label} of $\alpha$ is a pair $(\fr{i}, \fr{j})$ of the labels of $\partial \alpha$. The two points of $\partial E \cap L$ are also labeled according to the crossings at which $\partial E \cap L$ occurs. \begin{lem} For any arc $\alpha$ of $E \cap S^2$ with label $(\fr{i}, \fr{j})$, we have $i\neq j$. \end{lem} \noindent {\it Proof. } If both of the endpoints of $\alpha$ occur at the same crossing $\fr{i}$, we can observe that one of the two cases in Figure 4.2 occurs. In Figure 4.2 (a), $D$ is non-prime. In Figure 4.2 (b), there exists an arc $\alpha '$ of $E\cap S^2$ in $S^2 - {\rm Int}M$ such that the endpoints of $\alpha '$ occur at the same crossing $\fr{i}$, and that $\alpha '$ cuts off a disk $H$ from $S^2-\text{Int}M$ with $\text{Int}H\cap(E\cap S^2)=\emptyset$. We can surgery $E$ along $H$ so that we obtain two disks $E_1,\,E_2$ properly embedded in $S^3 - {\rm Int}M$. Since both endpoints of $\alpha '$ are in $R_{+}(\gamma)$ (or $R_{-}(\gamma)$), one of them, say $E_1$, intersects $L$ twice. Since $E$ is non-boundary-parallel, so is $E_1$ or $E_2$. If $E_2$ is, then it yields a compressing disk for $R$, a contradiction. Hence $E_1$ is a non-boundary-parallel $C$-product disk with $|E_1\cap S^2|<|E\cap S^2|$, a contradiction. \fpf \figbox{Figure 4.2} We look at an outermost disk $F \subset E$ (i.e., $F$ is the closure of a component of $E - S^2$ such that $F \cap S^2$ is connected). \begin{lem} Let $\alpha$ be an outermost arc of $E\cap S^2$ with label $(\fr{i}, \fr{j})$, cutting an outermost disk $F$ off $E$. Then we may assume that $i\neq j$ and that $i$ or $j=0$ if $F\cap L= \emptyset$ or (a point). \end{lem} \noindent {\it Proof. } By Lemma 4.1, we have $i \neq j$. Suppose $i \neq 0$ and $j \neq 0$. If $|F\cap L|=0$, $R$ is non-prime, a contradiction (Figure 4.3 (a)). If $|F\cap L|=1$, then either $|E \cap S^2|$ is not minimal, or $R$ is a plumbing (Figures 4.3 (b) and (c)). \fpf \figbox{Figure 4.3} Concerning outermost disks, we have two cases. {\bf Case B-1.} {\it There exists an outermost disk $F$ with $F \cap L =$ (a point).}\\ Let $\alpha$ be the arc $F \cap S^2\,(\subset E)$. By Lemma 4.2, we assume the label of $\alpha$ is $(\fr{0}, \fr{j})$, where $j \neq 0$. Let $\fr{k}$ be the label of the point of $\partial E \cap L$ on $F$. Then we have three cases; {\bf Subcase 1}: $k = 0$, {\bf Subcase 2}: $k = j$, and {\bf Subcase 3}: $k \neq 0$ and $k \neq j$. In Subcases 1 and 2, $D$ is non-prime (Figure 4.4 (a)). In Subcase 3, $R$ is a plumbing or we can isotope $E$ so that the outermost disk of Case B-1 is replaced by an outermost disk of Case B-2 (Figure 4.4 (b)). \figbox{Figure 4.4} \begin{lem} We may assume there exists no outermost disk of Case B-1. \end{lem} \noindent {\it Proof. } If the latter situation of Subcase 3 above occurs, we can view the above isotopy of $E$ as sliding a point of $\partial E \cap L$ out of $F$. Hence by repeating the above isotopies at most twice, we may eliminate outermost disks of Case B-1. \fpf {\bf Case B-2.} {\it There exists an outermost disk $F$ with $F \cap L = \emptyset$.}\\ By Lemma 4.2, we may assume $\alpha = F \cap S^2$ appears as in Figure 4.5. We note that outermost disks of this kind are typically found in the complementary sutured manifold for the fiber surface in Figure 3.2, which is obtained from a $2$-almost alternating diagram. The rest of the proof really depends on the almost-alternatingness of $D$. \figbox{Figure 4.5} \begin{lem} For any arc $\beta$ of $\partial E - (E \cap S^2)$, if $\beta \cap L = \emptyset$, then the endpoints of $\beta$ have different labels. \end{lem} \noindent {\it Proof. } Suppose the two endpoints of $\beta$ have the same label. Then $\beta$ appears as in Figure 4.6 and we can isotope $E$ to a $C$-product disk $E'$ such that $|E'\cap S^2|=|E\cap S^2|-1$, a contradiction. \fpf \figbox{Figure 4.6} \begin{lem} Suppose $E$ locally appears as in Figure 4.7 (a), i.e., $\fr{i},\ \fr{j}$ and $\fr{k}$ are the labels of points of $(\partial E\cap S^2) \cup (\partial E \cap L)$ sequential in $\partial E$ such that the former two points are connected by an outermost arc of $E\cap S^2$ and the last is a point of $\partial E \cap S^2$. Then $i, j, k$ are mutually different. \end{lem} \noindent {\it Proof. } By Lemma 4.2, we have $i\neq j$ and $i$ or $j=0$. Suppose $i=k$. Then we can find a compressing disk for $R$ in Figure 4.7 (b), a contradiction. By Lemma 4.4, we have $j\neq k$. \fpf \figbox{Figure 4.7} \begin{lem} We may assume that the following situation never occurs; The disk $E$ locally appears as in Figure 4.8 (a), i.e., $\fr{i},\,\fr{j},\,\fr{k}$ and $\fr{l}$ are the labels of points of $(\partial E\cap S^2) \cup (\partial E \cap L)$ sequential in $\partial E$ such that the former two points and the latter two are respectively connected by outermost arcs $\alpha_1$ and $\alpha_2$ of $E\cap S^2$. \end{lem} \noindent {\it Proof. } By Lemmas 4.2, 4.4 and 4.5, we may assume that $i=l=0$ and $j\neq k$. Then we obtain the conclusion, since in Figure 4.8 (b) $\alpha_2$ can not coexist with the arc of $\partial E-S^2$ connecting $\fr{k}$ and $\fr{l}$. \fpf \figbox{Figure 4.8} \begin{lem} Suppose $E$ locally appears as in Figure 4.9 (a), i.e., $\fr{i}, \fr{j}$ and $\fr{k}$ are the labels of points of $(\partial E \cap S^2) \cup (\partial E \cap L)$ sequential in $\partial E$ such that the former two points are connected by an outermost arc of $E \cap S^2$ and the third point is of $\partial E \cap L$. Then we may assume $i, j, k$ are mutually different. \end{lem} \noindent {\it Proof. } By Lemma 4.2, we have $i \neq j$, and $i$ or $j =0$. If $k=i$, then $R$ is compressible, a contradiction (see Figure 4.9 (b)). If $k=j$, we can reduce $|E\cap S^2|$ by isotopy, a contradiction (see Figure 4.9 (c)). \fpf \figbox{Figure 4.9} \begin{lem} Let $\fr{l}$ be the label of the point $x$ of $\partial E\cap L$. Suppose that the two points adjacent to $x$ in $\partial E$ are points of $\partial E\cap S^2$. Then the two adjacent points do not have the same label except for the case where they are both $\fr{l}$. \end{lem} \noindent {\it Proof. } Suppose the two points have the same label $\fr{i} (\neq \fr{l})$. By Lemma 4.1, we may assume that they are not connected by an arc of $E\cap S^2$. Then we can find a $C$-product disk $E'$ in Figure 4.10 such that $|E'\cap S^2|=0$, a contradiction. \fpf \figbox{Figure 4.10} Similarly we have the following lemma. \begin{lem} Suppose $E$ locally appears as in Figure 4.11(a), i.e., $\fr{i}, \fr{j}, \fr{k}$ and $\fr{l}$ are the labels of points of $(\partial E \cap S^2) \cup (\partial E \cap L)$ sequential in $\partial E$ such that the former two points are connected by an outermost arc of $E \cap S^2$, the third point is of $\partial E \cap L$ and that the fourth is of $\partial E \cap S^2$. If $k \neq l$, then $i, j, k$ and $l$ are mutually different. \end{lem} \noindent {\it Proof. } By Lemma 4.7, we may assume $i, j, k$ are mutually different. Then by Lemma 4.8, we have $l \neq j$. Suppose $l \neq k$ and $l = i$. Then by Lemma 4.2, $\fr{i} = \fr{l} = \fr{0}$ (Figure 4.11 (b)) or $\fr{j}=\fr{0}$ (c). In either case, we can find a $C$-product disk $E'$ such that $|E'\cap S^2|=0$, a contradiction. \fpf \figbox{Figure 4.11} \begin{lem} We may assume the following situation never occurs; The disk $E$ locally appears as in Figure 4.12 (a), i.e., $\fr{i}, \fr{j}, \fr{k}, \fr{l}$ and $\fr{m}$ are the labels of points of $(\partial E \cap S^2) \cup (\partial E \cap L)$ sequential in $\partial E$ such that the first two points and the last two points are respectively connected by an outermost arc of $E \cap S^2$, and that the third point is of $\partial E \cap L$. \end{lem} \noindent {\it Proof. } By Lemma 4.7, $k, l$ and $m$ are mutually different and hence by Lemma 4.9, $i, j, k$ and $l$ are mutually different. By Lemma 4.2, $i$ or $j=0$ and $l$ or $m=0$, and hence $m=0$, and by symmetry, we have $i=0$. Then $\fr{i}, \fr{j}, \fr{k}, \fr{l}$ and $\fr{m}$ appear as in Figure 4.12 (b), where we can find a $C$-product disk $E'$ such that $|E \cap S^2| > |E'\cap S^2|=1$, a contradiction. We note that $E'\cap L$ occurs at $\fr{k}$ and $\fr{l}$. \fpf \figbox{Figure 4.12} An arc $\varepsilon$ of $E \cap S^2$ is said to be {\it of level $2$} if it is not outermost and, for one component $E_1$ of $E - \varepsilon$, $E_1 \cap S^2$ is a union of outermost arcs in $E \cap S^2$. Suppose there is no arc of level $2$. Then by Lemmas 4.3, 4.6 and 4.10, we see that $E \cap S^2$ consists of only one arc $\alpha$ such that one component of $E - \alpha$ contains the two points of $\partial E \cap L$. Let $(\fr{0},\fr{j})$ be the label of $\alpha$, and let $\fr{k}$ and $\fr{l}$ be the labels of the two points of $\partial E\cap L$, where $\fr{0}, \fr{j}, \fr{k}$ and $\fr{l}$ appear in this order in $\partial E$. If $l=k$, then we can isotope $E$ so that $E \cap L = \emptyset$ and we have a compressing disk for $R$, for $E$ is not boundary parallel, a contradiction. Hence by Lemma 4.7, we can assume $j, k, l, 0$ are mutually different. In this case, $R$ desums into three surfaces $R_1, R_2$ and $R_3$ obtained by applying Seifert's algorithm to the almost alternating diagrams $D_1, D_2$ and $D_3$ respectively (Figure 4.13). \figbox{Figure 4.13} Hence we assume there is an arc of level $2$. Then by Lemmas 4.3, 4.6 and 4.10, we see that there exists an arc $\varepsilon$ of level $2$ such that one disk $E_{1}$ cut by $\varepsilon$ off $E$ contains one outermost arc of $E\cap S^2$ and satisfies one of the following conditions;\\ (*) $E_1\cap L=\emptyset$, \\ (**) $E_1\cap L=$ a point. If $E_{1}$ satisfies (*), by Lemmas 4.1 and 4.5, all four labels of points of $E_1\cap S^2$ are mutually different. Then, we can see that $D$ is non-prime or $R$ is a plumbing (Figure 4.14). \figbox{Figure 4.14} Thus we have: \begin{lem} We may assume that there is no arc of level $2$ which cuts a disk $E_{1}$ off E such that $E_{1}$ contains only one (outermost) arc of $E\cap S^2$ and that $E_{1}\cap L=\emptyset$. \end{lem} In what follows, we assume that there exists an arc $\varepsilon$ of level $2$ which cuts off $E$ a disk $E_{1}$ containing one outermost arc of $E\cap S^2$ and satisfying (**). By Lemma 4.3, we may suppose that $E_1$ appears as in Figure 4.15 (a) with labels $\fr{i},\,\fr{j},\,\fr{k},\,\fr{l}$ and $\fr{m}$. \begin{lem} All five labels in $E_{1}$ are mutually different. \end{lem} \noindent {\it Proof. } By Lemma 4.5, $i, j, k$ are mutually different. By Lemma 4.7, $l \neq k$ and $l \neq j$. We see $l \neq i$, for if not, $R$ appears as in Figure 4.15 (b) or (c), and in either case, $R$ is compressible, a contradiction. Now we have seen that $i, j, k, l$ are mutually different. Next suppose $m = l$. Then $R$ appears as in Figure 4.15 (d) or (e). In Figure 4.15 (d), $R$ is a plumbing or we can isotope $E$ to reduce $|E \cap S^2|$. In Figure 4.15 (e), $R$ is a Murasugi sum or we can isotope $R$ so that $D$ becomes an alternating diagram and the result follows from Theorem \ref{thm:fiber-hopf}. Hence we can assume $m \neq l$ and by Lemma 4.9, we see that $j, k, l, m$ are mutually different and by Lemma 4.1, $m \neq i$. \fpf \figbox{Figure 4.15} \begin{lem} We may assume $\fr{j}=\fr{0}$. \end{lem} \noindent {\it Proof. } If not, $\fr{k}=\fr{0}$ by Lemma 4.2. Then $R$ is a 6-Murasugi sum as in Figure 4.16. \fpf \figbox{Figure 4.16} \begin{lem} Let $\varepsilon$ and $E_1$ be as above. Then there is no arc $\varepsilon '$ of $E \cap S^2$ as in Figure 4.17 which cuts a disk $E_2$ off $E$ with the following conditions: \begin{enumerate} \item $E_{1}\subset E_{2}$, \item $({\rm Int }E_{2}-E_{1})\cap(E\cap S^2)=\emptyset$, \item $E_{2}\cap L=E_{1}\cap L= (1\,{\rm point})$. \end{enumerate} \end{lem} \figbox{Figure 4.17} \noindent {\it Proof. } By Lemmas 4.12 and 4.13, we may assume that $E_1$ appears as in Figure 4.18. Recall that $R$ is flat. Suppose that we have a disk $E_2$ as in Figure 4.17. Then the arc $\varepsilon'$ lies in some region of $S^2 - N(R)$. Hence, considering the orientation of $R$, we see that one of the following occurs;\\ (1) The point \textcircled{\small 1} is bounded by the same Seifert circle as one of the points \textcircled{\small 3} and \textcircled{\small 5}, \\ (2) The point \textcircled{\small 6} is bounded by the same Seifert circle as one of the points \textcircled{\small 2} and \textcircled{\small 4}.\\ In each case, we can find a $C$-product disk $E'$ such that $|E\cap S^2| > |E'\cap S^2| = 0 \ {\rm or\ } 1$, a contradiction. \fpf \figbox{Figure 4.18} \begin{lem}\label{lem:last} Let $E_1$ be as above. Then the following situation never occurs; The disk $E$ locally appears as in Figure 4.19, i.e., there is an outermost disk $F$ such that $\partial E - (E_1 \cup F)$ has a component $\beta$ which contains no point of $(\partial E\cap S^2) \cup (\partial E\cap L)$. \end{lem} \figbox{Figure 4.19} \noindent {\it Proof. } Suppose there exists such a disk $F$. Let $\alpha$ be an arc in $E\cap S^2$ which cuts $F$ off $E$, and $(\fr{s},\fr{t})$ the label of $\alpha$ where $\fr{s}$ is the label of an endpoint of $\beta$. First we examine the case where $E$ appears as in Figure 4.19 (a). If $s=i$ or $0$, we can find a $C$-product disk $E'$ such that $|E'\cap S^2|=1$, a contradiction (Figure 4.20 (a)). By Lemma 4.2, we have $s=0$ or $t=0$, and hence $t=0$. If $s = k$ or $l$, we can find a $C$-product disk $E'$ such that $|E' \cap S^2| = 0$, a contradiction (see Figure 4.20 (b)). By Lemma 4.4, we have $s \neq m$. Then we see that $R$ locally appears as in Figure 4.20 (c). It is impossible that $\partial E$ runs toward the dealternator $\fr{0}(=\fr{t})$ after passing through $\fr{s}$ because of the orientation of $R$. Second, we examine the case where $E$ locally appears as in Figure 4.19 (b). We can do this by the similar way to in the previous case. By Lemma 4.4, $s \neq i$. If $s = 0$ or $k$, we can find a $C$-product disk $E'$ such that $|E'\cap S^2|=0$, a contradiction. By Lemma 4.2, we have $s=0$ or $t=0$, and hence $t=0$. If $s = l$ or $m$, we can find a $C$-product disk $E'$ such that $|E'\cap S^2|=1$, a contradiction. Then we see that it is impossible that $\partial E$ runs toward the dealternator $\fr{0}(=\fr{t})$ before passing through $\fr{s}$. See Figure 4.21. \fpf \figbox{Figure 4.20} \figbox{Figure 4.21} Let $E_1' = E - E_1$. Then $E_1' \cap L$ is exactly one point, say, $x$. By Lemma 4.3, $E_1' \cap (E \cap S^2) \neq \emptyset$. By Lemmas 4.6 and 4.11, any arc of $E_1' \cap (E \cap S^2)$ which does not separate $\varepsilon$ and $x$ is outermost in $E_1'$. By Lemma 4.15, at least one of $E_1' \cap (E \cap S^2)$ separates $\varepsilon$ and $x$. Among such separating arcs, let $\alpha$ be the one closest to $\varepsilon$. Then by Lemma 4.15 again, the subdisk of $E$ between $\varepsilon$ and $\alpha$ contains no arc of $E_1' \cap (E \cap S^2)$. However, this contradicts Lemma 4.14. This completes the proof. \epf{\ref{thm:main}} \section{Proof of Theorem 1.5} In this section, we prove Theorem \ref{thm:hopf-plumbing}. Recall that a Seifert surface $R$ obtained by Seifert's algorithm is a union of {\it Seifert disks} and {\it Seifert bands}. \medskip \noindent {\bf Definition.} Let $R$ be a Seifert surface obtained by Seifert's algorithm. We say that two Seifert bands $B_1$ and $B_2$ of $R$ are {\it parallel} if they connect the same two Seifert disks. \medskip The following is a case where we can deplumb a Hopf band from a fiber surface: \begin{prop} \label{prop:para-deplumb} Let $R$ be a fiber surface obtained by Seifert's algorithm. Suppose $R$ has a pair of parallel bands $B_1$ and $B_2$. Then, we can deplumb a Hopf band from $R$. Moreover, we have the following;\\ (1) the parallel bands are of the same sign, and\\ (2) for each $i = 1, 2$, we can cut the band $B_i$ by deplumbing a Hopf band from $R$, i.e., $R$ is a plumbing of $R-B_{i}$ and a Hopf band. \end{prop} \noindent \noindent {\it Proof. } We denote by $L$ the link $\partial R$. We may assume that the Seifert circles, say, $C_1$ and $C_2$ connected by $B_1$ and $B_2$ bound mutually disjoint Seifert disks on the level 2-sphere $S^2$. First, suppose the pair of parallel bands are of the same sign. We may assume they appear as in Figure 5.1 (a). We explicitly show that $R$ is a plumbing of a Hopf band and the surface $R-B_{i}$. Move $L$ by isotopy as in Figure 5.1 (a) and let $R'$ be the surface as depicted. Apparently the Euler characteristic $\chi(R)$ is equal to $\chi(R')$. Hence by the uniqueness of fiber surfaces, we see that $R$ is isotopic to $R'$. Now we can deplumb a Hopf band from $R'$ as in Figure 5.1 (b). Then by retracing the above isotopy, we obtain the conclusion. Next suppose that the pair of parallel bands are of the opposite signs, i.e., that the twisting of $B_1$ is opposite. Then by the isotopy as implied by Figure 5.1 (a), we can find a compressing disk for $R'$, which contradicts the fact that fiber surfaces are of minimal genus and hence incompressible. \fpf \figbox{Figure 5.1} The following proposition assures that if a diagram $D$ has a Seifert circle $C$ which contains an alternating tangle diagram, then any Seifert surface obtained by applying Seifert's algorithm to $D$ has parallel bands. \begin{prop} \label{prop:alt-para} Suppose a Seifert surface $R$ obtained from an alternating diagram $D$ is a fiber surface. Then $R$ has parallel bands. Moreover, if $D$ is reduced, then for any band $B$ of $R$, there is a band $B'$ of $R$ which is parallel to $B$. \end{prop} \noindent {\it Proof. } By untwisting $R$ by isotopy if necessary, we may assume that $D$ is reduced. Moreover, we may assume that $D$ is unnested, because (1) by desumming along nested Seifert circles, we can decompose $R$ into fiber surfaces obtained from unnested alternating diagrams, and (2) if one of the decomposed surfaces has parallel bands, then so does $R$. Suppose a fiber surface $R$ for a link $L$ is obtained from a reduced unnested alternating diagram $D$. Then by \cite{Mu1960} (or \cite[Proposition 13.25]{BZ}), $L$ is a connected sum of $(2, n)$-torus knots or links. Moreover the arguments in \cite{Mu1960} shows that $D$ is the \lq standard' alternating diagram of a connected sum of $(2, n)$-torus knots or links. Hence we obtain the conclusion. \fpf \medskip \noindent {\it Proof of Theorem \ref{thm:hopf-plumbing}.} The \lq if' part follows from Theorem \ref{thm:fiber-sum}. We show the \lq only if' part, using Corollary \ref{cor:algorithm}, by induction on the first Betti number $\beta_1$ of $R$. If $\beta_1(R) = 1$, $R$ is a Hopf band, and hence the theorem holds. Assume the theorem holds for such surfaces with $\beta_1 < k$, and let $R$ be a Seifert surface with $\beta_1(R) = k$ obtained from an almost alternating diagram $D$. By untwisting $R$ if necessary, we may assume that $D$ is reduced. By Corollary \ref{cor:algorithm}, we know how $R$ decomposes into Hopf bands. Hence by the following four lemmas, we will see that we can deplumb a Hopf band from $R$, in such a way that by deplumbing a Hopf band, we cut a band of $R$ corresponding to an alternator. Therefore the deplumbed surface satisfies the assumption of induction so that we see that $R$ is a Hopf plumbing. \epf{\ref{thm:hopf-plumbing}} \begin{lem} \label{lem:claim1} If $R$ desums along a nested Seifert circle, then we can cut a band of $R$ by deplumbing a Hopf band from $R$. \end{lem} \noindent {\it Proof. } Suppose $D$ is nested, i.e., there exists a Seifert circle $C$ which contains another Seifert circle in both of its complementary regions in $S^2$. Then $R$ desums along $C$ into two surfaces, say, $R_1$ and $R_2$ such that $R_1$ is obtained from an alternating diagram and $R_2$ from an almost alternating diagram (cf. Proposition \ref{prop:nest}). Note that by Theorem \ref{thm:fiber-sum}, both $R_1$ and $R_2$ are fibers. By Proposition \ref{prop:alt-para}, we see that $R_1$ has parallel bands and hence so does $R$. Then by Proposition \ref{prop:para-deplumb}, we can cut a band of $R$ by deplumbing a Hopf band from $R$. \fpf \begin{lem}\label{lem:claim2} If $R$ is a connected sum, then we can cut a band of $R$ by deplumbing a Hopf band from $R$. \end{lem} \noindent {\it Proof. } Let $R$ be a connected sum of $R_1$ and $R_2$, where $R_1$ is obtained from an alternating diagram and $R_2$ from an almost alternating diagram by Proposition \ref{prop:prime}. Then by Theorem \ref{thm:fiber-sum} and Proposition \ref{prop:alt-para}, $R_1$ has parallel bands, which are also parallel in $R$, and hence, by Proposition \ref{prop:para-deplumb}, we can cut a band of $R$ by deplumbing a Hopf band from $R$. \fpf \begin{lem}\label{lem:claim3} If $R$ admits a decomposition of type (A), then we can cut a band of $R$ by deplumbing a Hopf band from $R$. \end{lem} \noindent {\it Proof. } Suppose $R$ admits a decomposition of type (A). Then we can deform $R$ to $R'$ by isotopy as depicted in Figure 5.2 (a), from which we can desum a fiber surface $R_1$ in Figure 5.2 (b). We can confirm that $R_1$ is obtained from an alternating diagram using Fact \ref{fact:tangle}. By Proposition \ref{prop:alt-para}, $R_1$ has parallel bands. Though $R'$ itself is not a surface obtained by Seifert's algorithm, we can apply the argument in the proof of Proposition \ref{prop:para-deplumb}, by regarding the inside of the dotted circle in Figure 5.2 (a) as a black box. Hence we can cut a band of $R'$ (which is a band in the image of $R_1$ in $R'$) by deplumbing a Hopf band from $R'$. This corresponds to cutting a band of $R$ by deplumbing a Hopf band from $R$. Note that we can confirm that the surface obtained from $R$ by this cutting the band satisfies the assumption of induction. \fpf \figbox{Figure 5.2} \begin{lem}\label{lem:claim4} If $R$ admits a decomposition of type (B), then we can cut a band of $R$ by deplumbing a Hopf band from $R$. \end{lem} \noindent {\it Proof. } According to whether the crossing visible in Figure 1.1 is an alternator or the dealternator, we have two cases. Let us call the former a decomposition of type (B1) and the latter of type (B2). Suppose that $R$ admits a decomposition of type (B1). Then by the same way as in the proof of Lemma \ref{lem:claim2}, we can cut a band of $R$ by deplumbing a Hopf band from $R$. Now assume $R$ does not admit a decomposition of type (B1). Then $R$ deplumbs into $R_1$ and $R_2$, which are both obtained from almost alternating diagrams (see Proposition \ref{prop:induction}). If $R_1$ or $R_2$ admits a decomposition of type (A), then we see, by the uniqueness of fiber surfaces, that $R$ also admits a decomposition of type (A), and the claim follows from Lemma \ref{lem:claim3}. Hence we assume that neither $R_1$ nor $R_2$ admits a decomposition of type (A). Inductively, if we can do a decomposition of type (A) or (B1) in the process of desumming $R$ into a union of Hopf bands, then we see that $R$ also admits a decomposition of type (A) or (B1). So we assume that $R$ desums into a union of Hopf bands using decompositions of type (B2) alone. Then by another inductive argument, we see that $R$ is a pretzel surface of type $(1, -3, \ldots, -3)$ or $(-1, 3, \ldots, 3)$. In this case, obviously we can cut a band of $R$ by deplumbing a Hopf band from $R$. \fpf {\bf Acknowledgment.} The authors would like to thank Professor Taizo Kanenobu, Professor Tsuyoshi Kobayashi, Professor Yasutaka Nakanishi and Professor Makoto Sakuma for their comments. Part of this work was carried out while the first author was visiting at University of California, Davis. He would like to express thanks to Professor Abigail Thompson and the department for their hospitality. \footnotesize{
\section{Introduction} Alcock \& Paczy\'nski (1979) suggested the possibility of using the clustering statistics of galaxies in redshift space to constrain the global geometry in the universe. The basic idea is that, since clusters of galaxies should not be preferentially aligned along any direction relative to a fixed observer, their average shape ought to be spherically symmetric. Therefore, if galaxies were following the Hubble expansion of the universe, without any peculiar velocities, the average extent of clusters in radial velocity $v_r$ (measured from redshifts) and their angular size $\psi$ are related to the physical size of the cluster $L$ by $v_r = H(z)\, L$, and $\psi = L/D(z)$, respectively. Here, $H(z)$ and $D(z)$ are the Hubble constant and the angular diameter distance at the redshift $z$ where the clusters are observed. The condition that clusters are spherical on average can then yield the value of $H(z)\cdot D(z)$. Of course, the effect of peculiar velocities must be included in order to apply this method, since any clustering induced by gravity will generally introduce peculiar velocities (Kaiser 1987) that will cause a distortion of similar or greater magnitude than the differences between cosmological models. Recently, the rate at which galaxies at high redshift are being identified has dramatically increased thanks to the Lyman limit technique, using the fact that the reddest objects among faint galaxies will often be galaxies at the redshift where the Lyman limit wavelength is between the two bands used to measure the color (Guhathakurta et al.\ 1990; Steidel \& Hamilton 1993; Steidel et al.\ 1996). For example, very red objects in $U-B$ are likely to be galaxies at redshift $z\simeq 3$. The galaxy correlation function, $\xi({\bf r})$, which measures the probability in excess of a random distribution of finding a galaxy at a real space separation vector ${\bf r}$ from another galaxy, has been measured for the first time for the population of Lyman-break galaxies (Giavalisco et al.\ 1998). The correlation length, defined to be the separation at which the excess probability is equal to that of a random distribution, has been estimated to be $\sim 2.1 h^{-1}$ Mpc (for an $\Omega_{0}=1$ universe; the symbol $\Omega$ is used here for the ratio of the density of matter in the universe to the critical density, the subscript $0$ indicates redshift zero), about half of the correlation length of galaxies at $z=0$. The bias, defined as the ratio of the correlation function of galaxies to that of matter at a fixed separation, is estimated to be large, $\sim 4$ for an $\Omega_{0}=1$ universe and smaller for universes with smaller dark matter content (Giavalisco et al.\ 1998). Count-in-cells analysis of the Lyman-break sample used in conjunction with a Press-Schechter mass function for the halos also indicate that these galaxies are likely to reside in rare, massive halos that existed at the time (Adelberger et al.\ 1998; Steidel et al.\ 1998; see also Coles et al.\ 1998 and Wechsler et al.\ 1998 for models of clustering of Lyman-break galaxies). These rare halos are expected to be much more clustered than the underlying matter distribution as originally suggested by Kaiser (1984) (see also Mo \& white 1996, for analytic models of bias as a function of the mass of halos). Both these analysis indicate that the population of Lyman-break galaxies is likely to be highly biased with respect to the underlying matter distribution. In this paper we investigate the feasibility of using the distortion of the redshift space correlation function of this population of galaxies to measure cosmological parameters. This possibility has been suggested before by Matsubara and Suto (1996) who proposed using the ratio of the value of the correlation function parallel to the line of sight to its value perpendicular to the line of sight at a fixed separation as a measure of the distortion. In this paper we express the angular dependence of the cosmological redshift space distortion of the correlation function as a multipole expansion. We are also specifically interested in applying this method to the highly biased, high redshift population of Lyman-break galaxies. Ballinger et al.\ (1996) have investigated the use of the full functional form of the redshift space power spectrum to separately measure the peculiar velocity effects and cosmological geometry effects. In essence this reduces to using both the quadrupolar as well as the octapolar distortion of the redshift space power spectrum to simultaneously constrain the cosmological constant as well as the parameter $\beta=\Omega^{0.6}/b$, where b is the linear theory bias. In this paper we fix the bias of the galaxy distribution by using the constraints on the matter power spectrum at redshift zero derived from observations of cluster abundances. On large scales the power spectrum of matter at any redshift is related to the power spectrum at redshift zero through the linear growth factor. We can then use the lowest order quadrupolar distortion of the power spectrum alone to constrain other cosmological parameters such as the cosmological constant. Ballinger et al.\ (1996) also estimated the errors involved in such a survey although in Fourier space. We estimate the errors in estimating cosmological parameters directly from the correlation function. On sufficiently large scales, where density fluctuations are in the linear regime, the angular form of the redshift space correlation function depends only on two parameters: the cosmological term $H(z)\cdot D(z)$, and the bias of the galaxy population. This paper presents a general method of estimating these two parameters from the basic data of a galaxy redshift survey, and evaluates the size of the survey that is necessary to determine the two parameters (or a combination of them, given other constraints from the galaxy distribution at the present time) with a given accuracy. We shall analyze the sensitivity of the method to a variety of cosmological models, placing special emphasis on models that contain a cosmological constant or a new component of the energy density of the universe with negative pressure christened Quintessence (e.g. Kodama \& Sasaki 1984; Peebles \& Ratra 1988; Caldwell et al.\ 1998), given the recent evidence from the luminosity distances to Type Ia supernovae (Garnavich et al.\ 1998; Perlmutter et al.\ 1997; Reiss et al.\ 1998) suggesting an accelerating universe. As pointed out by Alcock \& Paczy\'nski, the quantity $H(z)\cdot D(z)$ is more sensitive to this type of component than to space curvature. The paper is arranged as follows. In \S 2 we describe the effect of geometric distortion. In \S 3 we introduce the method for measuring the effects of cosmological geometry and peculiar velocity effects on the redshift space correlation function. In \S 4 we present predictions for a variety of cosmological models, and in \S 5 we estimate the errors in the observational determination of the redshift space correlation function contributed by shot noise and by the finite size of the observed volume. Our discussion is given in \S 6. \section{Method} A redshift survey consists of measuring the radial velocity and angular position of every galaxy included in the sample. We denote by ${\bf n}$ the unit vector along the line of sight, which, if the survey does not extend over a very large area, can be considered constant for all galaxies. Given a pair of galaxies, let $v$ be the difference between their radial velocities, and $\psi$ their angular separation. We define their vector separation in redshift space ${\bf w}$ as (see Figure 1) \begin{eqnarray} {\bf w} \cdot {\bf n} & = & v ~ , \nonumber \\ | {\bf w} - ({\bf w} \cdot {\bf n}) {\bf n} | & = & H(z)D(z) \, \psi ~ , \nonumber \\ w^2 & = & v^2 + \left[ H(z)D(z)\, \psi \right]^2 ~ . \end{eqnarray} where $H(z)$ and $D(z)$ are the Hubble constant and the angular diameter distance at the mean redshift of the survey, $z$. We also define $\mu$, for future use, as the cosine of the angle between the vector separation between two galaxies and the line of sight: \begin{equation} \mu=\frac{v}{w} \end{equation} The quantity $H(z)D(z)$ contains the dependence on the cosmological model. If we could measure the correlation function of galaxies directly in real space (measuring distances to galaxies instead of radial velocities), then the simple requirement that the correlation function should be isotropic would yield the value of $H(z)D(z)$. However, peculiar velocities should obviously introduce an anisotropy in the correlation function, and their effect needs to be included. \subsection{Model Dependence of $ H(z) D(z)$} Figures 2 and 3 show the ratio $ H(z) D(z) / H_{s}(z) D_{s}(z)$ for various models, where $ H_{s}(z) D_{s}(z)$ is the value of $ H(z) D(z)$ for a ``fiducial'' model, here adopted to be the Einstein-de Sitter model, with $\Omega_0=1$ in the form of pressureless matter. The symbol $\Omega_0$ is used here for the present ratio of the density of matter in the universe to the critical density. Two of the models shown in Figures 2 and 3 are the open model (with space curvature but no negative pressure components) and the cosmological constant (or $\Lambda$) model (with no space curvature and a component with pressure $p=-\rho c^2$). The third of the models shown is a Quintessence or Q model with no spatial curvature and a component with equation of state $p=-\rho c^2/3$. The quantity $ H(z) D(z)$ is much more sensitive to $\Lambda$ than to space curvature, and is also sensitive to the Q model, with a different redshift dependence. In general, a component of the energy density in the universe with negative pressure can have any equation of state, but the case $p=-\rho c^2/3$ implies an expansion mimicking exactly that of an open universe. Therefore, $H(z)$ in our Q model is exactly the same as in the open model. However, whereas in the open model the negative space curvature increases the angular diameter distance compared to the Einstein-de Sitter model, cancelling almost exactly the decrease in $H(z)$, the flat geometry of the Q model results in smaller angular diameter distances, so $ H(z) D(z)$ is smaller than in the Einstein-de Sitter model due to the decrease of $H(z)$. It is useful to note at this point that in order to obtain useful constraints on cosmological models, $ H(z) D(z)$ must be measured to an accuracy better than $\sim 10 \%$. In order to distinguish, between a cosmological constant and a Q model, $ H(z) D(z)$ must of course be measured at several redshifts with even higher accuracy. In practice, we can expect that any constraints obtained from measuring $ H(z) D(z)$ should be combined with other knowledge obtained, for example, from the luminosity distances to Type Ia supernovae. \section{Effect of peculiar velocities on the redshift space correlation function } For a given value of $ H(z) D(z)$ the effect of peculiar velocities on the shape of the redshift space correlation function is well described in the literature (e.g. McGill 1990; Hamilton 1992;Fisher 1995) and the redshift space correlation function, $\tilde{\xi}({\bf w})$, is given by : \begin{eqnarray} \tilde{\xi}({\bf w})&=&\sum_{l=0,2,4} D_{l} (\beta,w,z) \cdot P_{l}(\mu) \, , \label{eqnxi} \end{eqnarray} where, \begin{eqnarray} \beta \simeq~\frac{\Omega(z)^{0.6}}{b(z)}, \nonumber \end{eqnarray} where $b(z)$ is the bias parameter for the class of objects under survey and $\Omega(z)$ is the ratio of the density of matter to the critical density at redshift $z$. The coefficients of the expansion in Legendre polynomials, $D_{l}$, can be expressed as: \begin{eqnarray} D_{l} (\beta,w,z) &=& (-1)^{l} \cdot A_{l} (\beta) \cdot \xi_{l}(w,z) ~, \end{eqnarray} where \begin{eqnarray} A_{0}&=&\left({1+ \frac{2}{3} \beta + \frac{1}{5} \beta^{2}} \right)~, \nonumber \\ A_{2}&=&\left( {\frac{4}{3} \beta - \frac{4}{7} \beta^{2} } \right)~, \nonumber \\ A_{4}&=& \left( {\frac{8}{35} \beta^{2}} \right)~, \nonumber \label{eqnAl} \end{eqnarray} and \begin{eqnarray} \xi_{l}(w,z)&=& \frac{b(z)^{2}}{2 \pi^2} \int dk~k^{2}~P(k,z)~j_{l}(kw)~, \end{eqnarray} and $j_{l}$ is the lth order spherical Bessel function. The function $P(k,z)$ is the linear matter power spectrum at redshift z in terms of the k vector in velocity space. Note that $\xi_{0}(w,z)$ is proportional to the real space matter correlation function at redshift z. Hence, $D_{0}$ is equal to the real space two point correlation function for this class of objects except for the factor of $A_{0}[\beta(z)]$. We also mention here that the $D_{2}$ coefficient is negative implying a squashing of the contours of the correlation function along the line of sight as is expected due to the peculiar velocities from infall on large scales. Throughout this paper we use the simple model of linear, local, constant biasing of galaxies, i.e. the overdensity in the number of galaxies, $\delta_{g}(\vec{w})$, is given by $b \times \delta_{m}(\vec{w})$, where $\delta_{m}(\vec{w})$ is the overdensity in matter and b the bias. For general deterministic local bias models, this is valid in linear theory where $\delta_{g} < 1$ ( for $b>1$ and $\delta_{m} \sim -1, \, \delta_{g} < -1$ is unphysical) (Gazta\~naga \& Baugh 1998). Hence, our results are likely to be valid on large scales where the correlation function is smaller than one. In reality, biasing is not easily modeled since it depends on the complex process of galaxy formation, which is poorly understood. Several alternative models of galaxy biasing have been suggested, including non-local biasing mechanisms (Babul \& White 1991; Bower et al.\ 1993) and stochastic biasing (Dekel \& Lahav 1998; Tegmark \& Peebles 1998). However, for stochastic (local) models, on large scales, the bias (the ratio of the correlation function of galaxies to that of matter) will be independent of scale (Scherrer and Weinberg 1998) as in the case of a linear, local, constant biasing scheme, although the variance in the measured correlation function will be larger for such models. On the other hand, non-local models of galaxy biasing in which the efficiency of galaxy formation is modulated coherently over large scales, result in scale dependent bias. In the absence of a well motivated model for bias, we have assumed the simplest scale independent model for the bias. It is valid only on large scales and is not generally valid for non-local biasing models. We also mention that we have only taken the linear infall velocities into account in calculating the redshift space correlation function (see Equation \ref{eqnxi}). On small scales non-linear velocity effects (`Fingers of God') will also be important (e.g. see Fisher et al.\ 1994 for the redshift space correlation function of IRAS galaxies). \subsection{Effect of geometric distortion} In order to test the magnitude of the geometric distortion, we calculate the anisotropy introduced in the correlation function by varying $ H(z) D(z)$ about its fiducial value, $ H_{s}(z) D_{s}(z)$. Let the product $ H(z) D(z)$ for any other model be given by \begin{equation} H(z) D(z) = H_{s}(z) D_{s}(z) \cdot \sqrt{1+\alpha(z)}~, \end{equation} where $\alpha(z)$ is defined as the geometric distortion parameter. Then, using equations (1) and (2) we have, \begin{eqnarray} w^{2} &=& w_{s}^{2} ~\eta ^{2} (\alpha,\mu_{s}) ~, \nonumber \\ \mu^{2} &=&\frac{\mu_{s}^{2}}{\eta ^{2}(\alpha,\mu_{s})} ~, \end{eqnarray} where \begin{equation} \eta^{2} (\alpha,\mu_{s}) = 1 + \alpha (1 - \mu_{s}^{2}) ~. \end{equation} We can now express equation (\ref{eqnxi}) in terms of the variables $w_{s}, \mu_{s}$ in the fiducial model: \begin{equation} \tilde{\xi}({\bf w}) = \sum_{l=0,2,4} D_{l}(\beta,w_{s} \cdot \eta,z) \cdot P_{l}(\frac{\mu_{s}}{\eta})~. \end{equation} Rewriting this as a series in $P_{l}(\mu_{s})$, \begin{equation} \tilde{\xi}({\bf w}) =\sum_{l} C_{l} (\beta,w_{s},z) \cdot P_{l}(\mu_{s})~, \end{equation} one can immediately see from the angular dependence in $\eta$ that the expansion is an infinite series in $P_{l}(\mu_{s})$, with the new coefficients of the Legendre polynomials, $C_{l}$, being given by, \begin{equation} C_{n}(\beta,w_{s},z)=\left( {\frac{2n+1}{2}} \right) \sum_{l=0,2,4} \int D_{l} (\beta,w_{s} \cdot \eta,z) \cdot P_{l}(\frac{\mu_{s}}{\eta}) \cdot P_{n}(\mu_s)~d\mu_{s}~. \end{equation} Thus, expressing the coefficients $D_{l}$ of a given model in terms of fiducial coordinates introduces angular distortion in the redshift space correlation function. \section{Results for the geometric distortion} In this section we present our results for the sensitivity of the anisotropy of the correlation function to the geometric distortion parameter, $\alpha$. We consider here a galaxy survey with a mean redshift of $3$, the typical redshift of the current Lyman limit galaxy surveys. Our fiducial model is $\Omega_{0}=1.0$, with the standard cold dark matter (SCDM) power spectrum. On large scales the power spectrum at redshift $3$ is related to the power spectrum at redshift zero by the linear growth factor. We adopt the cluster normalization for the power spectrum at redshift zero, obtained by requiring that the observed density of galaxy clusters with a given X-ray temperature matches the theoretical prediction. The constraint obtained in this way can be expressed in terms of the fluctuation in a sphere of radius $8h^{-1}$ Mpc, $\sigma_{8}$, given by (Eke et al.\ 1996): \begin{eqnarray} \sigma_{8}&=&0.52~\Omega_{0}^{-0.46 \Omega_{0}}\,,\,{\rm for}\, \Lambda_{0}=0 ~,\nonumber \end{eqnarray} and \begin{eqnarray} \sigma_{8}&=&0.52~\Omega_{0}^{-0.52 \Omega_{0}}\,,\,{\rm for}~\Omega_{0}+\Lambda_{0}=1 . \end{eqnarray} In Figure 4 we plot the $C_{l}(\beta,w_{s})$ coefficients for $l=0,2$ and $4$, for our fiducial model (lighter lines) and the $\Lambda$ model with $\Omega_{0}=0.3$, $\Lambda_{0}=0.7$ (bold lines). The horizontal axis has been labeled both in units of velocity ($w_{s}$) and comoving space separation $s$, calculated for the fiducial model. The observed correlation function is given by the monopole term in the Legendre polynomial expansion, $C_{0}(\beta,w_{s})$. In order to match the value of the $C_{0}$ coefficient to unity at the observed correlation length of $2.1 h^{-1} \, {\rm Mpc}$ (comoving) for $\Omega_{0}=1$ (Giavalisco et al.\ 1998), which corresponds to a correlation velocity of $450 \, {\rm km\, s}^{-1}$, the bias required is $b=4$. For the $\Omega_{0}=0.3$, $\Lambda_{0}=0.7$ model, the bias required to match the computed $C_{0}$ coefficient to $1$ at the observed correlation velocity of $450 \, {\rm km\, s}^{-1}$ is $2.3$. On account of the large bias in these two models, we can rely on the linear theory that we have used for peculiar velocity distortions of the correlation function we have shown in \S 3 for $C_{0} \lesssim 1$, or $w_{s} \gtrsim 400 \, {\rm km\, s}^{-1}$. Our goal is to measure the multipoles $C_{l}(\beta,w_{s})$ of the correlation function and use it to constrain the geometric distortion parameter, $\alpha$. From Figure 4 we see that on scales of approximately $10^{3} \, {\rm km\, s}^{-1}$, the $C_{2}$ coefficient is a $10 \%$ perturbation on the monopole term, whereas the octapolar term $C_{4}$ is a smaller contribution at $\sim 3 \%$ for the $\Lambda_{0}=0.7$ model. Once we fix the bias, the $C_{0}$ coefficients for the two models shown are similar to each other, except on very large scales. The quadrupolar coefficients for the two models on the other hand are very different. The $C_{2}$ coefficient for the $\Lambda$ model, affected by geometric distortion, is larger by a factor of 2 compared to the $\Omega_{0}=1$ model and comes to within a factor of 2 of the monopole term on scales $\sim 10^{4} \, {\rm km\, s}^{-1}$. As we mentioned in \S 2, the $C_{2}$ coefficient is less than zero which implies a squashing of the contours of constant $ \tilde{\xi}({\bf w}) $ along the line of sight. Thus we see that for our choice of the fiducial model, the primary effect of geometric distortion caused by a model with a positive cosmological constant is to cause a further squashing of the contours of $ \tilde{\xi}({\bf w}) $. We mention here that the $C_{4}$ coefficient is even more sensitive to geometric distortion than the $C_{2}$ coefficient. Its value is approximately $10$ times larger for the $\Lambda_{0}=0.7$ model as compared to the fiducial $\Lambda=0$ model. A measurement of the $C_{4}$ coefficient will give us additional information with which to test the bias model that we have used. It will be interesting to compare the value of the linear bias parameter parameter derived from a simultaneous measurement of the cosmological constant and the $\beta$ parameter using both the $C_{4}$ and $C_{2}$ coefficients, to the value obtained by comparison of the galaxy distribution to the matter power spectrum at redshift zero. We now show that the difference in angular distortion of the correlation function in the two models is primarily due to the change in the distortion parameter and is not strongly dependent on the choice of the power spectrum. In Figure 5, we plot the coefficients $C_{l}$, for the fixed cosmological model $\Omega_{0}=0.3,\Lambda_{0}=0.7$, but two different correlation functions. The bold lines correspond to the power spectrum of the $\Lambda$ model with these same parameters. The lighter lines are for the same cosmological model, but with the power spectrum of an $\Omega_{0}=1$ CDM model as a function of $k/H(z)$. We see from this figure that at velocity separations $ \gtrsim 10^4 \, {\rm km\, s}^{-1}$, the differences in the power spectra dominate the differences in the $C_{l}$ coefficients. But on smaller scales the geometric distortion effect is the most important effect. In particular the $C_{2}$ and $C_{4}$ coefficients are similar once the monopoles for both models are normalized to unity at the observed correlation length. Thus the ratios of the coefficients $C_{2}/C_{0}$ and $C_{4}/C_{0}$ are only weakly dependent on the shape of the power spectrum. This shows that it should be possible to measure the geometric distortion parameter even if the power spectrum is not known accurately from independent methods. \section{Error estimates} In this section we compute the accuracy in the measurement of the multipoles of the redshift space correlation function from a typical survey volume and test the feasibility of the method described above. Currently, the typical observed fields have a size $\sim 12 \arcmin$ on each side. The redshift range of each field extends from $z=2.6$ to $z=3.4$ with a surface density of approximately $1.25$ Lyman-break objects per square arc minute within this redshift range (Adelberger et al.\ 1998). In our fiducial model ($\Omega_{0}=1$), this corresponds to a width of $ 2 \times 10^{3} \, {\rm km\, s}^{-1}$ and a depth of $ 6 \times 10^{4} \, {\rm km\, s}^{-1}$. We consider for the purpose of error estimation, a wide field of view of $\sim 3^{\circ}$. We shall later discuss the scaling of the errors with the angular size of the field of view. Any detailed calculation of the errors in a survey will depend upon the precise geometry of the survey volume and the selection effects involved in the survey. Here, we consider two of the sources of error, shot noise and cosmic variance. Shot noise is caused by the discrete nature of the galaxies from which we measure the correlation function. Cosmic variance arises due to the finite volume we use to estimate a statistical quantity. We calculate these errors for a single cylindrical survey volume with a radius of $ 1.5^{\circ}$ ($75 h^{-1} \, {\rm Mpc}$ for $\Omega_{0}=1$ model), and depth extending from $z=2.6$ to $z=3.4$ ($300 h^{-1} \, {\rm Mpc}$ for $\Omega_{0}=1$ model). At the current estimate of surface density of Lyman-break galaxies of $1.25$ per square arcmin (Adelberger et al.\ 1998), approximately $30000$ galaxies would be included in our survey volume. \subsection{Shot noise} In order to estimate the redshift space correlation function, we bin pairs of galaxies with respect to their separation velocity $w_{s}$ (computed in the fiducial model) in widths of $\Delta w_{s}$. The redshift correlation function is then estimated (denoted by subscript E) as, \begin{equation} \tilde{\xi}_{E}({\bf w}_{s})=\frac{N_{p}(w_{s},\mu_{s})}{\overline{N}_{p}(w_{s},\mu_{s})}-1, \label{eqnrxi} \end{equation} where $N_{p}(w_{s},\mu_{s})$ are the number of pairs with separations between $w_{s}$ and $\Delta w_{s}$ with the separation vector making an angle ${\rm cos}^{-1}(\mu_{s})$ with the line-of-sight and $\overline{N}_{p}(w_{s})$ is the ensemble average of a random distribution of the same quantity. There are various different estimators for the correlation function discussed in the literature that minimize the error of the estimator due to the unknown true average density of the galaxies at the redshift of the survey (for a discussion see Hamilton 1993). Our shot noise will be dominated by the small number of pairs of galaxies we have in each of our bins. Since we are currently only interested in an estimate of this error, we have adopted the simpler estimator for the correlation function. To analyze the data from a survey one should use a more sophisticated estimator to minimize its variance. Using equation (\ref{eqnrxi}) we obtain the estimate of the $C_{l}$ coefficients as given below for $l \neq 0$ \begin{equation} C_{l,E}(w_{s})=\frac{2l+1}{2} \frac{1}{\overline{N}_{p}(w_{s})} \cdot \sum_{i=1}^{N_{p}} P_{l}(\mu_{si}) \,, \label{eqnClp} \end{equation} where $N_{p}$ is the number of pairs with separations between $w_{s}- \Delta w_{s}$ and $w_{s}+ \Delta w_{s}$, and $\overline{N}_{p}(w_{s},\mu_{s})$ is the average number of pairs for a random distribution of galaxies in the same bin . The summation is performed over the pairs of galaxies (denoted by subscript i) in the bin centered at $w_{s}$. In order to calculate the statistical average of the estimator, we have to perform two integrals. First, for a given number of pairs separated by $w_{s}$, we average over their possible orientations. The probability that a given pair of galaxies with separation $w_{s}$ is oriented along $\mu_{s}$ is given by $\psi(w_{s},\mu_{s})$, where \begin{equation} \psi(w_{s},\mu_{s}) d\mu_{s}=\frac{1+ \tilde{\xi}(w_{s},\mu_{s})}{1+C_{0}(w_{s})} d\mu_{s}. \end{equation} The $1+C_{0}(w_{s})$ factor in the denominator comes from normalizing $1+ \tilde{ \xi}(w_{s},\mu_{s})$ over $\mu_{s}$. Here we have assumed that a given pair of galaxies can have any orientation with respect to the line of sight. This is clearly not true, for example for a pair of galaxies close to the edge of the survey volume. In order the circumvent this difficulty we consider a smaller volume within the total survey volume which we call the ``reduced volume'', hereafter denoted as $V_{R}$ such that the edges of $V_{R}$ are a distance $w_{s}$ away from the edges of the total survey volume. We only consider pairs of galaxies such that at least one of the galaxies is within $V_{R}$ . For a random distribution of galaxies, a pair chosen in this way is not biased to be aligned along a particular direction. We can see that the largest separation at which we can measure the coefficients $C_{l}$ is the radius of the survey for which $V_{R}$ goes to zero. Secondly, we have to average over the distribution of the number of pairs of galaxies in each bin. Calculating the averages (denoted by brackets) yields \begin{equation} <C_{l,E}(w_{s})>= \frac{2l+1}{2} \frac{<N_{p}(w_{s})>}{\overline{N}_{p}(w_{s})} \int \psi(w_{s},\mu_{s}) P_{l}(\mu_{s}) d \mu_{s} \,. \end{equation} This gives us that $<C_{l,E}(w_{s})> = C_{l}(w_{s})$. This result can also be shown to hold for the monopole term. In a similar way to the calculation of the statistical average of the $C_{l}$ coefficients, we can calculate the mean square variation of $C_{l}$ coefficients. Using equation (\ref{eqnClp}) we have, \begin{equation} C_{l,E}^{2}(w_{s})=\left( \frac{2l+1}{2} \right)^{2}\left( \frac{1}{\overline{N}_{p}} \right)^{2} \left( \sum_{i=1}^{N_{p}} P_{l}(\mu_{si})^{2} + 2 \sum_{i<j}^{N_{p}} P_{l}(\mu_{si})P_{l}(\mu_{sj}) \right). \label{Clsqpoiss} \end{equation} The statistical average of the above equation gives the mean square variance of the $C_{l}$ coefficients, $<C_{l,E}^{2} - <C_{l,E}^{2}>>$, denoted by $\sigma_{l}^{2}$, as, \begin{equation} \sigma_{l}^{2}(w_{s})=\left( \frac{2l+1}{2} \right) \frac{(1+C_{0}(w_{s})) }{\overline{N}_{p}(w_{s})}. \end{equation} We mention here that in deriving the above equation we have assumed a Poisson distribution for the number of pairs in each bin. This assumption is not strictly valid since every pair separation is not independent. Hence, one may expect some underestimation in the Poisson errors we have calculated but this should be small since the second term in equation (\ref{Clsqpoiss}) is proportional to $C_{l}^{2}$. Figure 6 shows the expected $1 \sigma$ error for the $C_{l}$ coefficients due to shot noise. Each successive bin is centered at $w_{s}$ with value $1.5$ times that of the previous bin and hence, each bin has width $2/5 w_{s}$. The average number of pairs $\overline{N}_{p}(w_{s})$ in the bin centered at $w_{s}$ and width $2 \times \Delta w_{s}$ for a random distribution of galaxies within the survey volume is given by \begin{equation} \overline{N}_{p}(w_{s})=\frac{\overline{n}_{g}^{2}}{2} \times {\rm V_{R}} \times 4 \pi w_{s}^{2} 2 \Delta w_{s} \,, \end{equation} where $\overline{n}_{g}$ is the average density of galaxies within the survey volume. This is an underestimate of the number of pairs in the bin since it counts only half of the pairs of galaxies of which one of the galaxies is outside $V_{R}$. At larger separations this underestimation is maximum since, in this case, a larger fraction of all the pairs in the bin have one of the galaxies outside of $V_{R}$. Thus our shot noise is an overestimate by a factor $\leq \sqrt{2}$. We can see from Figure $6$ that with shot noise alone, the $C_{l}$ coefficients are best measured in the velocity range $10^{2} \lesssim w_{s} \lesssim10^{4} \, {\rm km\, s}^{-1}$ for a survey of the size and geometry that we have assumed. The errors on the multipoles scale as $(2l+1)^{\frac{1}{2}}$, and so they are smaller for the $C_{0}$ coefficient and higher for the $C_{4}$ coefficient as compared to the quadrupole. For scales close to the radius of the survey the shot noise error increases rapidly since $V_{R}$ is now very small. For scales $\sim 10^{3} \, {\rm km\, s}^{-1}$, with shot noise alone, we can measure the $C_{2}$ coefficient to a few percent accuracy, both for the fiducial model as well as the $\Lambda$ model and hence distinguish a large cosmological constant as in our model with $\Lambda_{0} = 0.7$ to high statistical significance. The shot noise error on the $C_{4}$ coefficient is small for the $\Lambda_{0}=0.7$ model we have shown but larger for models with smaller cosmological constants. Considering shot noise alone, on scales of $\sim\, 10^{3} \, {\rm km\, s}^{-1}$, the $C_{4}$ coefficient can be measured if it is present at the level of a few percent of the monopole, which in turn would indicate a large energy density in the form of a cosmological constant or some form of quintessence. We also note here that the number of pairs of galaxies at a fixed separation is proportional to $V_{R}$. Hence for separations small compared to the radius of the survey, the shot noise error scales as the inverse of the angular size of the survey. \subsection{Cosmic variance} The cosmic variance of a survey volume results from the sparse sampling of the universe made by the small survey volume. It occurs even if the overdensity at each point within the survey volume is accurately known, and is independent of the number of observed galaxies. We estimate the cosmic variance in this section using linear theory. A finite volume estimate (denoted by subscript E) of the correlation function is given by, \begin{equation} \tilde{\xi}_{E}(\vec{w_{s}},\hat{n}) = \frac{1}{{\rm V_{R}}} \int_{{\rm V_{R}}} d^{3}x \, \delta (\vec{x},\hat{n}) \delta (\vec{x} + \vec{w_{s}},\hat{n}) \, . \label{eqnxiE} \end{equation} In the above equation, $\vec{x}$ is constrained to be within $V_{R}$ such that its boundary are a distance $w_{s}$ away from that of the full survey volume and $\vec{x} + \vec{w_{s}}$ is within the full survey volume. For every point $\vec{x}$ within $V_{R}$, the overdensities at $\vec{x}$ and $\vec{x}+\vec{w_{s}}$ are accurately known. The ensemble average of the estimator gives, \begin{eqnarray} < \tilde{\xi}_{E}(\vec{s})> = \frac{1}{{\rm V_{R}}} \int_{{\rm V_{R}}} d^{3}x \, \tilde{\xi}(\vec{s}) \, , \nonumber \\ =\tilde{\xi}(\vec{s}) \, , \end{eqnarray} where as previously, quantities without the subscript $E$ stand for their true values. Similarly, \begin{equation} <C_{l,E}(w_{s})>=C_{l}(w_{s}). \end{equation} The variance in $C_{l,E}(w_{s})$ can be computed using, \begin{equation} <C_{l,E}^{2}(w_{s})>=\left( \frac{2l+1}{2} \right)^{2} \int d\mu_{s1} \int d\mu_{s2} < \tilde{\xi}_{E}(s,\mu_{s1}) \tilde{\xi}_{E}(s,\mu_{s2}) > P_{l}(\mu_{s1}) P_{l}(\mu_{s2}) \, , \label{eqnCl} \end{equation} where, \begin{equation} < \tilde{\xi}_{E}(w_{s},\mu_{s1}) \tilde{\xi}_{E}(w_{s},\mu_{s2}) > =\frac{1}{{(\rm V_{R}})^{2}} \int d^{3}x_{1} \int d^{3}x_{2} < \delta (\vec{x_{1}}) \delta (\vec{x_{1}}+\vec{w_{s1}}) \delta (\vec{x_{2}}) \delta (\vec{x_{2}}+\vec{w_{s2}})> \,, \end{equation} where $|\vec{w}_{s1}|=|\vec{w}_{s2}|$ and $\hat{w}_{s1} \cdot \hat{n}= \mu_{s1}$, $\hat{w}_{s2} \cdot \hat{n}= \mu_{s2}$. In order to simplify the above expression, we approximate the overdensities to be in the linear regime. The linear overdensities are Gaussian distributed and the four point expression in the above equation can be expressed in terms of two point correlation functions : \begin{eqnarray} <\delta (\vec{x_{1}}) \delta (\vec{x}_{1}+\vec{w}_{w1}) \delta (\vec{x}_{2}) \delta (\vec{x}_{2}+ \vec{w}_{s2}) > = \tilde{\xi} ( \vec{w}_{s1}) \tilde{\xi} ( \vec{w}_{s2}) \nonumber \\ + \tilde{\xi} ( \vec{x}_{1}- \vec{x}_{2}) \tilde{\xi} ( \vec{x}_{1}- \vec{x}_{2}+ \vec{w}_{s1}- \vec{w}_{s2}) \nonumber \\ + \tilde{\xi} ( \vec{x}_{1}- \vec{x}_{2}+ \vec{w}_{s1}) \tilde{\xi} ( \vec{x}_{1}- \vec{x}_{2}- \vec{w}_{s2}). \label{eqn4pt} \end{eqnarray} Thus the last two terms in equation (\ref{eqn4pt}) contribute to the root mean square variance in $C_{l,E}(w_{s})$. Since $\tilde{\xi}_{E}(w_{s},\mu_{s})$ is independent of the azimuthal angle in equation (\ref{eqnCl}), we can also integrate over this angle. Therefore we can express the root mean square variance as, \begin{eqnarray} <\sigma_{C_{l,E}}^{2}(w_{s})>=\left(\frac{2l+1}{2 {\rm V_{R}}} \right)^{2} \int d^{3}x_{1} \int d^{3}x_{2} \frac{d \Omega_{1}}{2 \pi} \frac{ d \Omega_{2}}{2 \pi} P_{l}(\mu_{s1}) P_{l}(\mu_{s2}) \nonumber \\ \left\{ \tilde{\xi} ( \vec{x}_{1}- \vec{x}_{2}+ \vec{w}_{s1}) \tilde{\xi} ( \vec{x}_{1}- \vec{x}_{2}- \vec{w}_{s2}) + \tilde{\xi} ( \vec{x_{1}}- \vec{x_{2}}) \tilde{\xi} ( \vec{x}_{1}- \vec{x}_{2}+ \vec{w}_{s1}- \vec{w}_{s2}) \right\}. \label{eqnsigma} \end{eqnarray} The method we employed in the calculation of the above integrals is detailed in the Appendix. Figure 7 displays the expected cosmic variance errors for our survey volume for the multipole coefficients. For all three coefficients, the cosmic variance dominates the error on large scales, while the shot noise contribution is larger on smaller scales. The error is smallest in the region $w_{s} \sim 3000 \, {\rm km\, s}^{-1}$, so this is the best scale at which to measure the quadrupole and octapole coefficients and hence estimate the geometric distortion factor. As mentioned before, the cosmic variance error that we have calculated assumes linear theory and hence we have underestimated the contribution to the errors from fluctuations and non-linear velocity effects on small scales. As mentioned in \S 3 if we adopt a local but stochastic model for the distribution of galaxy number density as a function of the underlying mass density, then the bias will still be scale independent on large scales, but there will be a larger variance in the measured correlation function. In the absence of a well motivated stochastic biasing model, we have not estimated the variance in the $C_{l}$ coefficients arising from such a model of the bias. Depending on the true nature of bias, we may be underestimating the variance of the measured correlation function. A more precise estimate of the error can only be given by the direct analysis of numerical simulations, a project we plan to return to in a later paper. We note here that we have not assumed that the mean overdensity within the survey volume is zero. The fluctuation in the mean overdensity is the primary source of the cosmic variance error for the monopole component of the correlation function. This fluctuation of course does not affect the higher multipoles of the correlation function and hence on small scales the error on the higher multipoles is smaller than on the monopole coefficient. With the combined shot noise and cosmic variance errors the $C_{2}$ coefficient can be measured to a few percent accuracy both for the fiducial $\Omega_{0}=1$ model as well as the $\Lambda$ models which have a larger quadrupole coefficient compared to the fiducial model. Thus, with our estimate of the errors, we can distinguish a geometric distortion factor of about $15 \%$ corresponding to a $\Lambda$ model with $\Lambda_{0}=0.7$ to high statistical significance. From Figures 6 and 7 we also see that for our survey volume, the $C_{4}$ coefficient can be measured to $\sim 20 \%$ accuracy for the $\Lambda_{0}=0.7$ Model. The errors are larger for models with smaller cosmological constants. Therefore, a measurement of the octapolar coefficient is possible if it is present at the level of a few percent of the monopole on scales of $\sim 3\times 10^{3} \, {\rm km\, s}^{-1}$ as in case of a large cosmological constant. Since the error on $C_{4}$ is large, a simultaneous measurement of both the $\beta$ parameter as well as the cosmological constant from the anisotropy of the redshift space correlation function alone is difficult. This has been indicated earlier by Ballinger et al.\ (1996). If the the octapolar coefficient can be measured, and the bias parameter constrained, it will be interesting to compare its value to the one obtained by comparison of galaxy clustering to the assumed underlying matter distribution. But we emphasize that when we assume that the amplitude of the matter power spectrum is known, and only one parameter needs to be measured from the redshift space correlation function, then the quadrupolar geometric distortion effect of the cosmological constant can be measured to high accuracy. For scales much smaller than the radius of the survey, the cosmic variance error scales as the inverse square root of the volume of the survey and hence as the inverse of the angular size of the survey. Thus both shot noise and cosmic variance have similar dependence on the angular size of the survey on small length scales. A large cosmological constant may be distinguished with high statistical significance for smaller angular size surveys depending upon other sources of error. Considering only the shot noise and the cosmic variance that we have estimated, for a survey of angular size $1^{\circ}$, with a factor of three increase in the errors, we can still measure the quadrupolar coefficient affected by a geometric distortion parameter of $15 \%$ with an accuracy of approximately $10 \%$ on a scale of $3 \times 10^{3}\, {\rm km\, s}^{-1}$. The error is larger for smaller distortion factors. Since a variation of the cosmological constant from zero to $0.7$ changes the quadrupolar coefficient by a factor of 2, we can use a linear relation between the two to make an approximate estimate of the accuracy with which the value of the cosmological constant can be measured. This gives us that a large cosmological constant, for which the error in the difference of the $C_{2}$ coefficient with respect to its value in the fiducial model is small, can be constrained with an error bar of approximately $20 \%$ with a $1^{\circ}$ field of view. Since in fact this linear relation is incorrect and the geometric distortion parameter is more sensitive to a variation in the cosmological constant when it is large (Ballinger et al.\ 1996), the error we have quoted will be somewhat smaller for large $\Lambda_{0}$ ($\gtrsim 0.5$). For a field of view of this size, the $C_{4}$ coefficient can also be measured, although with a large error of $\sim 60 \%$, if it is present at the level of a few percent of the monopole as in the case of geometric distortion with respect to the fiducial $\Omega_{0}=1$ model by a cosmological constant $\Lambda_{0}=0.7$. For a smaller field of view, the monopole coefficients have to be measured on scales smaller than $3000 \, {\rm km\, s}^{-1}$ where shot noise is the dominant source of error. For example for a field size of $1/2^{\circ}$, the quadrupole coefficient corresponding to a $15 \% $ geometric distortion parameter can still be measured to an accuracy of approximately $50 \% $ on a scale of $10^{3} \, {\rm km\, s}^{-1}$. Hence it can be distinguished from the fiducial $\Omega_{0}=1$ model at the $2 \sigma$ level. For smaller scales the error is larger while to measure the distortion parameter at larger scales a larger field size is required. Thus a field at least $1/2^{\circ}$ in diameter, corresponding to an area approximately four times the currently used field size, is required to distinguish a $\Lambda_{0}=0.7$ model from our fiducial $\Omega_{0}=1$ model. \section{Discussion and Conclusions} In this paper we have investigated the feasibility of using the high redshift population of Lyman-break galaxies to measure the geometric distortion effect and hence constrain cosmological parameters. The method is particularly sensitive to components of energy density with negative pressure and in particular to the cosmological constant. The principal advantage of using this population of galaxies is their high bias with respect to the underlying matter distribution. This tends to suppress the peculiar velocity effects and makes it easier to measure the geometric distortion effect. As pointed out by Ballinger et al.\ (1996), a simultaneous measurement of the bias and the cosmological constant using the redshift space distortion alone is difficult except in case of a large cosmological constant. In this paper we assumed that the matter power spectrum at redshift $3$ is related by the linear growth factor to the matter power spectrum at redshift zero which is constrained by observations of cluster abundances. We fixed the bias of the Lyman-break galaxies by comparing their clustering to the assumed matter power spectrum at redshift $3$. Then we only need to measure one parameter, the geometric distortion parameter, from the anisotropy of the correlation function. This permits us to use the lowest order quadrupolar distortion of the redshift space power spectrum to constrain the geometric distortion parameter to high accuracy. In cases of a large energy density in a cosmological constant or quintessence, the octapolar coefficient may also be measured. An interesting test would then be to compare the value of the bias parameter derived from the additional information provided by the octapolar term to that determined by comparing the galaxy clustering to the matter power spectrum. We estimated that in order to distinguish a flat model with $\Lambda_{0}=0.7$ from the Einstein-de Sitter case, at least a $1/2^{\circ}$ sized circular field of view is required. Currently the observation fields have sizes of approximately $10^{'}$, which are too small for measurements of geometric distortion, both due to shot noise and cosmic variance. It is preferable to measure the distortion effect on large scales where the effects of peculiar velocities can be analytically computed using linear theory. For this reason, it is better to use a single large field of view than to combine data from several small fields of view which provide data only on smaller scales. For a more accurate measurement of the distortion parameter larger field sizes are required. We estimated that for a field size of $3^{\circ}$, the best scale at which to measure the ratio of the quadrupole coefficient to the monopole is approximately $3000 \, {\rm km\, s}^{-1}$, or $15 h^{-1} \, {\rm Mpc}$ in the $\Omega_{0}=1$ model and somewhat smaller for a smaller field. Since the difference in the quadrupolar coefficients for the flat $\Lambda_{0}=0$ and $\Lambda_{0}=0.7$ models can be measured to $\sim 20 \%$ accuracy with a circular field of diameter $1^{\circ}$, we made a rough estimate that a large cosmological constant $\gtrsim 0.5$ can be measured with this precision. Our cosmic variance was estimated using the linear correlation function and we have underestimated the error due to fluctuations and non linear velocity effects on small scales. We have also used a very simple local non-stochastic scale independent model for the bias. Stochastic bias will lead to variance in the measured correlation function which we have not accounted for. A full calculation of the errors including non linear effects will require analysis of numerical simulations, which we will discuss in a future paper. \acknowledgements I wish to acknowledge Jordi Miralda-Escud\'e, my thesis advisor, who gave me the original motivation for this work and for the numerous insightful comments and discussions I have had with him. I also wish to thank Patrick MacDonald, Brian Mason and David Moroz for their comments on the paper. I would also like to acknowledge the anonymous referee for his comments and suggestions that have improved the content and presentation of the paper. \newpage
\section{Introduction} For a variety of systems the interplay between nonlinearity and disorder results in novel and fascinating phenomena \cite{j:1,j:2}. Particularly, the study of soliton dynamics in inhomogeneous and disordered media has received a great deal of attention in recent years \cite {j:3,j:4,j:5,j:6,j:7,j:8,j:9,j:10,j:11,j:12,j:13,j:14} since it concerns real condensed matter systems and phenomena. It is well known that transport properties in inhomogeneous and disordered media can change dramatically when the nonlinearity allows the creation of solitons. As a first step in the study of soliton dynamics in disordered media many authors explored the interaction between a soliton and an isolated impurity. This soliton-impurity interaction has been modeled by a point-like impurity and a structureless soliton. In real situations we can have several inhomogeneities of different kinds. When the distance between impurities is considerable higher than the width of solitons and impurities the traditional approach gives a correct result. When this is not so, we witness a series of surprising phenomena. The dimensionless Klein-Gordon-like equations model a wide variety of soliton bearing systems\cite {j:1,j:2,j:11,hanggi,jj:1,jj:2,jj:3,jj:4,jj:5,jj:6,jj:7,jj:8,jj:9,jj:10,jj:15,jj:16,jj:20,j:40,j:41,j:50}% , including charge density waves, Josephson junctions, structural phase transitions, crystal growth, polymers, proton conductivity, macromolecules and hydrogen-bond chains: \begin{equation} \phi _{xx}-\phi _{tt}-\frac{dU}{d\phi }={\cal F}(\phi ,\phi _t,x,t); \label{equ_1} \end{equation} here $U=U(\phi )$ is a potential that possesses at least two minima\cite {j:61}, meanwhile ${\cal F}(\phi ,\phi _t,x,t)$ represent additional forces (external forcing, dissipation, presence of impurities, inhomogeneous external fields, coupling to other degrees of freedom). The soliton solution of Eq. (\ref{equ_1}) is usually treated as a structureless point-like particle. A richer dynamics is unveiled when the extended character of the soliton is taken into account. For instance, in a previous work \cite{j:6}, Gonz\'alez and Ho\l yst studied the $\phi ^4$ equation ($U=\frac 18(\phi ^2-1)^2$): \begin{equation} \phi _{xx}-\phi _{tt}-\gamma \phi _t+\frac 12\left( \phi -\phi ^3\right) =-F(x)-G(x,t). \label{e:1} \end{equation} Particularly, they showed that the zeroes of $F(x)$ (for $G\equiv 0$) play the roll of equilibrium positions for the soliton. In the case that only one zero, $x_0$ exists, the condition for stability for the kink/antikink is: \begin{equation} \left[ \frac{dF}{dx}\right] _{x=x_0}\left\{ \begin{array}{ll} >0 & \mbox{for the kink.} \\ <0 & \mbox{for the antikink.} \end{array} \right. \label{con:1} \end{equation} For the stable case, the inhomogeneity can trap the soliton, but in the case that $F(x)$ possesses more than one zero, the stability condition may become too complex. Here the extended character of the soliton arises, and interesting phenomena appear such as the interaction between the structure of the soliton (which is not a point-like particle) and the inhomogeneities, and between the shape modes of the soliton themselves. The external force $% F(x)$ can change the spectrum of small oscillations about the soliton and additional bounded states can exist. For the soliton of the $\phi ^4$ equation without perturbation, there are only two bounded states (besides the continuous spectrum): the translational mode and the shape mode. For specific values of the parameters that define the force $F(x)$ used in Ref. \cite{j:6}, not only an increase of the number of shape modes can exist, but in certain cases these can be unstable. Moreover the continuous spectrum can lose stability and the soliton becomes unstable against interaction with phonons. The above mentioned authors also considered the problem with the time dependent force $G(x,t)$ and found that if $G(x,t)$ has a spatial shape such that it coincides with one of the eigenfunctions of the stability operator of the soliton, then it is possible to get resonance if the frequency of the force also coincides with the resonant frequency of the considered mode. This means, for example, that energy can be given to the translational mode (or any of the shape modes) using a $G(x,t)$ coupled to the translational mode (or any of the shape modes). When $F(x)$ has three zeroes, this is equivalent to a double well potential (like the one found in the Duffing equation\cite{j:6,j:15}) chaotic motion of the soliton is possible applying an additional periodic force $G(x,t)$ for a determined set of values of the parameters. In this paper we take into account the extended character of both the soliton and the impurity. We show that these considerations lead to the existence of a finite number of soliton internal modes that underlies a rich spatiotemporal dynamics. We introduce impurities of the $N(x)\phi $ type, where $N(x)$ is a function with a bell shape. An impurity of this kind but using delta functions has been presented in Ref. \cite{j:14}. In our paper we consider a finite width impurity and a finite width kink and show the striking differences between this approach and the traditional one (structureless solitons and delta-function-like impurities). We present a model for which the exact stationary soliton solution in the presence of inhomogeneities can be obtained and the stability problem can be solved exactly. To achieve this purpose we solve an inverse problem in order to have external perturbations which are generic and topologically equivalent to well-known bifurcation model systems\cite{j:51}. We choose the ``exact'' solution such that the differential operator that appears in the stability problem is a Posch-Teller potential that can be solved exactly. Besides, the ``generated'' external force and the impurities have important physical properties. In particular, the inhomogeneous force $% F(x)$ is equivalent to the pitchfork bifurcation canonical form\cite{j:15} and the $N(x)\phi $ impurity is topologically equivalent to the $\delta (x)\phi $ type impurity that is very frequently used. Furthermore, we demonstrate the sensibility of the soliton internal dynamics to the inhomogeneity width even for the isolated impurity case. Our paper is organized as follows. In Sec. II we present a description of our model and we give specific physical interpretations of the equations under consideration. In Sec. III we study the equilibrium positions of the soliton and its stability. We analyze the interaction of the soliton with the inhomogeneity created by the interplay between a finite-width and $\phi $% -dependent impurity and the already studied inhomogeneities independent of $% \phi $ \cite{j:6}. We also consider the action of time dependent forces fitted to the shape of the translational mode of the soliton. In Sec. IV we describe the numerical simulations that confirm the theoretical results. We use the Karhunen-Lo\`eve decomposition to relate the excitation of the shape modes spectrum of the soliton with the bifurcations. In Sec. V we present the interaction of the soliton with radiation modes. Finally, in Sec. VI we summarize and discuss our results and also present some concluding remarks. In the Appendices we outline the numerical method and present the Karhunen-Lo\`eve decomposition. \section{The model} The topological solitons studied in the present paper possess important applications in condensed matter physics. For instance, in solid state physics, they describe domain walls in ferromagnets or ferroelectric materials, dislocations in crystals, charge-density waves, interphase boundaries in metal alloys, fluxons in long Josephson junctions and Josephson transmission lines, etc.\cite{j:11,jj:7} Although some of the above mentioned systems are described by the $\phi ^4$% -model and others by the sine-Gordon equation (and these equations, in their unperturbed versions, present differences like the fact that the sine-Gordon equation is completely integrable whereas the $\phi ^4$-model is not) the properties of the solitons supported by sine-Gordon and $\phi ^4$ equations are very similar. In fact, these equations are {\it topologically equivalent} and very often the result obtained for one of them can be applied to the other\cite{jj:7}. Here we consider the $\phi ^4$ equation in the presence of inhomogeneities and damping: \begin{equation} \phi _{xx}-\phi _{tt}-\gamma \phi _t+\frac 12\left( \phi -\phi ^3\right) =-N(x)\phi -F(x), \label{a:1} \end{equation} where $F(x)$ is a function with (at least) one zero and $N(x)$ is a bell-shaped function that rapidly decays to zero for $x\rightarrow \pm \infty $. In ferroelectric materials $\phi $ is the displacement of the ions from their equilibrium position in the lattice, $\frac 12\left( \phi -\phi ^3\right) $ is the force due to the anharmonic crystalline potential, $F(x)$ is an applied electric field, and $N(x)$ describes an impurity in one of the anharmonic oscillators of the lattice\cite{s:3}. In Josephson junctions, $% \phi $ is the phase difference of the superconducting electrons across the junction, $F(x)$ is the external current, and $N(x)$ can describe a microshort or a microresistor\cite{j:5}. In a Josephson transmission line it is possible to apply nonuniformly distributed current sources ($F(x)$) and to create inhomogeneities of type $N(x)$ using different electronic circuits in some specific elements of the chain\cite{j:11,s:5}. In the present paper the functions $F(x)$ and $N(x)$ will be defined as, \begin{equation} F(x)=\frac 12A(A^2-1)\tanh (Bx), \label{a:10} \end{equation} \begin{equation} \ N(x)=\frac 12\frac{(4B^2-A^2)}{\cosh {}^2(Bx)}. \label{a:11} \end{equation} The case $F=const.$ has been studied in many papers (see e.g. \cite{jj:7}). Here Eq.~(\ref{a:10}) represents an external field (or a source current in a Josephson junction) that is almost constant in most part of the chain but changes its sign in $x=0$ (this is very important in order to have soliton pinning\cite{j:6}). Microshorts, microresistors and/or impurities in atomic chains\cite{j:5} are usually described by Dirac's delta functions ($\delta (x)$) where the width of the impurity is neglected. The function $N(x)$ is topologically equivalent to a $\delta (x)$ but it allows us to consider the influence of the width of the impurity. \section{Stability Analysis} \label{sec2} Let us consider the Eq. (\ref{a:1}) and assume the existence of a static kink solution $\phi _k(x)$ that corresponds to a soliton placed in a stable equilibrium state created by the inhomogeneities $F(x)$ and $N(x)$. Analyzing the small amplitude oscillations around the kink solution $\phi _k(x)$, \begin{equation} \label{a:2} \phi (x,t)=\phi _k(x)+\psi (x,t), \end{equation} we get, for the function $\psi (x,t)$, the following equation: \begin{equation} \psi _{xx}-\psi _{tt}-\gamma \psi _t+\frac 12(1-3\phi _k^2+2N(x))\psi =0. \label{a:3} \end{equation} Studying of the stability of the equilibrium solution $\phi _k(x)$ leads to the following eigenvalue problem (we have introduced $\psi (x,t)=f(x)\exp (\lambda t)$ into Eq.~(\ref{a:3})): \begin{equation} -f_{xx}+\frac 12(3\phi _k^2-1-2N(x))f=\Gamma f, \label{a:4} \end{equation} where $\Gamma \equiv -\lambda ^2-\gamma \lambda $. Let us now study some particular cases: If $F(x)\equiv 0$, and the function $% N(x)$ is described by the expression (\ref{a:11}), then it can be shown that the exact solution for the kink in equilibrium at the position $x=0$ is, \begin{equation} \label{a:6} \phi _k(x)=\tanh (Bx), \end{equation} and that the discrete eigenvalue spectrum is described by the following formula, \begin{equation} \label{a:7} \Gamma _n=B^2(\Lambda +2\Lambda n-n^2-2); \end{equation} where the parameter $\Lambda $ is defined as, \begin{equation} \label{a:8} \Lambda (\Lambda +1)=\frac 1{B^2}+2. \end{equation} The integer part of $\Lambda $ ($\left[ \Lambda \right] $) defines the number of modes of the discrete spectrum. From Eq.~(\ref{a:7}) the stability condition for the translational mode ($% n=0 $) can be obtained: \begin{equation} \label{a:9} 4B^2<1. \end{equation} It is worth noticing that if the coefficient of $N(x)$ is negative, the equilibrium position created by the impurity of the $N(x)\phi $ type is stable for the soliton, and that the stability condition is independent of the polarity of the soliton. The opposite occurs when a inhomogeneity like $% F(x)$ is considered. Furthermore, it is necessary to point out the differences between the case in which the soliton equilibrium position is due to a zero of $F(x)$, and the case in which the kink is trapped in the effective well created by the impurity $N(x)\phi$. In the former the characteristic width and the number of internal modes that can be excited are smaller than those for the free kink, and even smaller than those for the kink in equilibrium in an unstable position. While, in the latter the characteristic width and the number of internal modes that can be excited are greater than those for the free kink and for the kink in unstable equilibrium in a repulsive impurity. This is due to the inverse proportionality between the width of the impurity and $B$ (just like the width of the kink). The impurity is more stable as $B$ diminishes. Finally, let us consider the case in which both types of inhomogeneities ($% F(x)$ and $N(x)\phi $) are present in Eq.~(\ref{a:1}). These functions are defined as in equations (\ref{a:10}) and (\ref{a:11}). For these functions the exact solution describing the static soliton can be written: \begin{equation} \label{a:12} \phi _k(x)=A\tanh (Bx). \end{equation} The spectral problem (Eq.~(\ref{a:4})) brings the following eigenvalues for the discrete spectrum: \begin{equation} \label{a:13} \Gamma _n=\frac 12A^2-\frac 12+B^2(\Lambda +2\Lambda n-n^2-2); \end{equation} here $\Lambda $ is defined as, \begin{equation} \label{a:14} \Lambda (\Lambda +1)=\frac{A^2}{B^2}+2. \end{equation} The stability condition for the translational mode is, \begin{equation} 16B^4+2B^2(5-7A^2)+(1-A^2)^2<0. \label{a:15} \end{equation} When this condition is not fulfilled (the equilibrium position $x=0$ is unstable) and $A^2>1$, then there will exist three equilibrium points for the soliton: two stable (at points $x=x_1>0$ and $x=x_2<0$) and one unstable at point $x=0$. This happens because for large values of $\left| x\right| $ the leading inhomogeneity is $F(x)$, which is non-local and not zero at infinity. This inhomogeneity acts as a restoring force that pushes the soliton towards the point $x=0$. As a result of the competition between the local instability induced by $N(x)\phi$ at $x=0$ and the non-local inhomogeneity $F(x)$, an effective double-well potential is created. This is equivalent to a pitchfork bifurcation. Also note that when $A=0$ then $\Lambda =1$, i.e., there is an unstable mode created by the impurity $N(x)\phi $ even for a flat initial condition. This contrasts with the $\Lambda =0$ result expected for a delta-like impurity and with the result of Gonz\'alez and Ho\l yst\cite{j:6} for the other kind of inhomogeneity ($F(x)$). \section{Quasiperiodic and chaotic solitons} In this section we present numerical results for Eq. (\ref{equ_1}) perturbed with inhomogeneous external forces, impurities and time-periodic forces. First, we show, for inhomogeneities of the $F(x)$ type, the bifurcations leading to the chaotic regime. Later, we study quasiperiodic regimes as the soliton internal modes are excited by an impurity of the type $N(x)\phi $. Finally, we present bifurcations as the internal modes are excited for the case in which the two types of inhomogeneities ($F(x)$ and $N(x)\phi $) are present. \subsection{Scenario of inhomogeneous external forces} We simulate the following equation, \begin{equation} \phi _{xx}-\phi _{tt}-\gamma \phi _t+\frac 12\left( \phi -\phi ^3\right) =-F(x)-G(x,t), \label{b:1} \end{equation} where static and time-dependent forces are inhomogeneous and given by the following expressions: \begin{equation} F(x)=\frac 12A(A^2-1)\tanh (Bx)+\frac 12A(4B^2-A^2)\sinh (Bx)\cosh {}^{-3}(Bx) \label{b:2} \end{equation} and \begin{equation} G(x,t)=\nu \cos (\omega t)\left( \frac 1{\cosh {}^2(E(x-x_1))}+\frac 1{\cosh {}^2(E(x+x_1))}\right). \label{b:3} \end{equation} The force $F(x)$, firstly introduced in Ref. \cite{j:6}, can be obtained as a result of an inverse problem for which an effective potential equivalent to the canonical form of the pitchfork bifurcation is desired. For $A^2>1$ and $B^2\left( \sqrt{1+\frac{6A^2}{B^2}}-1\right) <1$, we are in presence of a double-well potential {\it for the soliton}, (i.e., the model is a Duffing-like\cite{j:15} soliton oscillator). Note that a force with three zeroes does not imply necessarily the existence of a double-well potential for the soliton{\it .} We will attend to the dynamics of the center of mass as defined by Eq. (\ref {aa:3}). Figure~\ref{Fig.1} presents the bifurcation diagram ($A=\sqrt{\frac 32}$, $B=\sqrt{\frac 1{10}}$, $\Lambda =2$, $\omega =1.22$, $E=0.1$, $% x_1=2.5 $ and $\gamma =0.450$) as the amplitude $\nu $ of the time-dependent driving force is increased. The period doubling cascade for low values of the control parameter corresponds to oscillations of the soliton in one well. At $\nu =0.312$ an unusual discontinuous transition interrupts the cascade: the system switches from a period-eight solution to a period-two solution but in the other well as can be appreciated in the time series presented in figures \ref{Fig.2}(a)-(b). At $\nu =0.318$ the period-eight solution is recovered and a two-band regime with n-periodic (Fig. \ref{Fig.3}% (a)), quasiperiodic (as the two-torus attractor presented in Fig. \ref{Fig.3}% (b)) and chaotic attractors. The quasiperiodic attractor is a spatiotemporal effect: it is generated by the activation of a shape mode of the soliton which provides a frequency that is incommensurate to the driving frequency. Figures \ref{Fig.1} and \ref{Fig.2}(c) reveals for $\nu =0.332$ the onset of jumps of the soliton between the two wells. This regime corresponds to Duffing-like chaos for the soliton. Figure \ref{Fig.3}(b) also presents the Poincar\'e map which reveals the high-dimensional chaotic motion of the soliton which can be ascribed to an increased activation of shape modes (which is not possible for a Duffing-like chaotic particle). The intense activity around $x_{c.m.}=0 $ (Fig. \ref{Fig.1}) is also due to the extended character of the soliton. We have verified that at $\nu =0.356$ the soliton prefers its deformation instead of its destruction and its dynamics returns to a periodic solution (figures \ref{Fig.1} and \ref{Fig.2}(d)). Figures \ref{Fig.4}(a)-(b) present time series for the center of mass of chaotic solitons. In the first case ($\nu =0.3$ and $\gamma =0.505$, the rest parameters are preserved) the soliton jumps between two wells. When the damping is increased (Fig. \ref{Fig.4}(b), $\gamma =0.550$), the soliton is constrained to move mainly in one well. Corresponding Poincar\'e maps (Fig. \ref{Fig.5}(a)-(b)) show the contrast of dimensionality (and consequently, of the number of modes and/or of the effective number of degrees of freedom) of these chaotic motions. Figures \ref{Fig.5}(c)-(d) present the spatiotemporal evolution for these chaotic solitons. Figure \ref{Fig.5}(c) evidences the jumps of the domain wall between the two wells, whereas Fig. \ref{Fig.5}(d) shows such a domain wall oscillating in one of the wells. Notwithstanding, note also in Fig. \ref{Fig.5}(d) the appearance and disappearance of deformations (depicted by yellow zones) in the other well due to tunnelling of mass/energy. \subsection{Scenario of finite-width impurity} We simulate the following model, \begin{equation} \phi _{xx}-\phi _{tt}-\gamma \phi _t+\frac 12\left( \phi -\phi ^3\right) =-N(x)\phi -G(x,t), \label{b:4} \end{equation} where external forces are time-dependent and inhomogeneous. Right-hand functions are given by, \begin{equation} N(x)=\frac{\frac 12\left( 4B^2-1\right) }{\cosh {}^2(Bx)} \label{b:5} \end{equation} and \begin{equation} G(x,t)=\frac{\nu \cos (\omega t)}{\cosh {}^\Lambda (Bx)}. \label{b:6} \end{equation} One may think that soliton dynamics in presence of an attractive impurity must be simple. Notwithstanding, that is not the case: as has been discussed in Section~\ref{sec2} when the width of the impurity is finite, a large number of internal modes can be excited. It can exist energy exchange between these modes and the translational mode bringing a complex dynamics. Figure \ref{Fig.6}(a) shows the phase space for a period-three solution ($% A=1.0$, $B=0.25$, $\Lambda =3.772$, $\omega =1.0$, $\nu =0.16$ and $\gamma =0.1$). As the amplitude $\nu $ of the time-dependent driving force is increased, the excitation of an internal mode can provide a frequency incommensurate to the driving frequency, generating a two-torus quasiperiodic attractor (Fig. \ref{Fig.6}(b) for $\nu =0.195$, the rest parameters are the same as the previous case). On the other hand, decreasing of the damping parameter $\gamma $ can provide even more incommensurate frequencies to the driving frequency as reveals the three-torus quasiperiodic motion presented in figures \ref{Fig.6}(c)-(d) for $\gamma =0.01$ and $\upsilon =0.12$. \subsection{Scenario of finite-width impurity and inhomogeneous forces} We will consider the following model: \begin{equation} \phi _{xx}-\phi _{tt}-\gamma \phi _t+\frac 12\left( \phi -\phi ^3\right) =-N(x)\phi -F(x)-G(x,t), \label{b:7} \end{equation} where $N(x)$ and $F(x)$ are given by Eq.~(\ref{a:10}) and Eq.~(\ref{a:11}), and $G(x,t)$ is given by the following expressions, \begin{equation} G(x,t)=\nu \cos (\omega t)\left( \frac 1{\cosh {}^\Lambda (B(x-x_1))}+\frac 1% {\cosh {}^\Lambda (B(x+x_1))}\right). \label{b:9} \end{equation} We will study the sequence of bifurcations as the driving amplitude $\nu $ is increased and other parameters remain fixed ($A=1.22$, $B=0.32$, $\Lambda =2$, $\omega =1.22$, $x_1=2.5$ and $\gamma =0.3)$. Figure \ref{Fig.7} presents the time-averaged spatial profile for different values of the amplitude $\nu $ of the time-dependent driving force: $\nu =0.20$ (period one), $\nu =0.28...0.45$ (quasiperiodicity), $\nu =0.55$ (chaos) and $\nu =0.60$ (period one). Figures \ref{Fig.8}(a)-(d) present a sequence of ``quasiperiodic bifurcations'': torus entangles as the amplitude of the time-dependent driving force increases and shape modes are activated. Figures \ref{Fig.9}(a)-(b) present an chaotic soliton for $\nu =0.55$. The Poincar\'e map (Fig. \ref{Fig.9}(a)) reveals the high-dimensional chaotic motion due to the activation of many internal modes whereas Fig. \ref{Fig.9}% (a) presents the temporal evolution of the soliton and shows that the kink profile is still sustained. For these parameters the soliton is at the edge of its destruction due to the activation of the shape modes. We can consider this regime as fully developed shape chaos for the soliton. Along this paper we had emphasized the importance of the internal modes (or shape modes) of the soliton that can be excited as the soliton interacts with inhomogeneities. Therefore we perform a Karhunen-Lo\`eve decomposition for the sequence of dynamic attractors already presented for the scenario of finite impurity and homogeneous force. Figure \ref{Fig.10}(a) reveals the increasing excitation of the discrete internal modes as the system evolves into a chaotic regime as well as the sudden change of the spectra for the final state that correspond to solution in which periodic motion is regained ($\nu =0.60$). The periodic solution for $\nu =0.60$ corresponds to the highest deformity of the kink profile (Fig. \ref{Fig.7}). This agrees with the higher contribution to the dynamics of the first two modes whereas all the rest of the modes decreased their contribution. Furthermore, the first shape mode replaces the translational mode as the leading mode of the dynamics. Figure \ref{Fig.10}(b) presents the leading Karhunen-Lo\`eve eigenmodes for the period-one solutions that initiates and ends the sequence of bifurcations considered in this section. The eigenmode for $\nu =0.20$ appears to be the superposition of a pair of translational modes centered at the equilibrium points for the soliton. Similar situation occurs for $\nu =0.60$ but the eigenvalue appears to be the superposition of a pair of shape modes. Figures \ref{Fig.11}(a)-(b) the striking difference of the temporal evolution of the period-one solitons for $\nu =0.20$ and $\nu =0.60$. \section{Interaction of the Soliton with Radiative Modes} Any process that involves inelastic interactions or accelerations of the soliton leads to the emission of quasi-linear waves (radiation). This phenomenon occurs by means of the modes of the continuous spectrum (radiation modes\cite{j:40}) that for the case of the unperturbed $\phi ^4$% -equation are given by \begin{equation} f_k(x)=e^{ikx}\left[ 3\tanh ^2\left( \frac x2\right) -6ik\tanh \left( \frac x% 2\right) -\left( 1+4k^2\right) \right] . \label{cc:1} \end{equation} The interaction of a soliton with an inhomogeneity results in an emission of radiation that can be calculated using the method of McLaughlin and Scott% \cite{j:40} that relies on the construction of a Green's function that consists of a bilinear combination of eigenfunctions (i.e., in our case, of the eigenfunctions that we have already presented). The radiation problem also has been addressed using other perturbative methods\cite{j:5}. In this paper we have focused our attention on the stability of the translational mode. Note that when the soliton is in an equilibrium position created by the impurities that is stable for the translational mode, the soliton as a whole is stable against the emission or absortion of radiation (these modes are usually called phonon modes). Moreover, in this case the kink is oscillating in an effective potential well for which radiative effects are exponentially small\cite{j:5}. Notwithstanding, we remark that when the equilibrium position is unstable the shape modes can also become unstable (under certain conditions) and this can destroy the soliton. Moreover, under a certain condition the modes of the continuous spectrum can also be unstable. This is very surprising as it leads the soliton to be unstable against the emission and the absortion of radiation modes. This interesting effect is shown in Figure \ref{Fig.12} where we present the interaction of a radiation mode with the soliton under instability conditions for the continuous spectrum. The usual methods for calculating the radiation can not give this result. For instance, the phenomenon shown in Figure \ref{Fig.12} is produced by the evolution of the system \begin{equation} \phi _{xx}-\phi _{tt}-\gamma \phi _t+\frac 12\left( \phi -\phi ^3\right) =-F(x), \label{cc:2} \end{equation} \begin{equation} F(x)=B(4B^2-1)\tanh (Bx) \label{cc:3} \end{equation} under initial conditions that represent the supperposition of a radiation mode and a soliton, \begin{equation} \phi (x,0)=2B\tanh (Bx)+C\left\{ 3\sin (Bx)\tanh (Bx)-6\cos (Bx)\tanh (Bx)-5\sin (Bx)\right\} , \label{cc:4} \end{equation} \begin{equation} \phi _t(x,0)=0 \label{cc:5} \end{equation} For $4B^2-1<0$ the translational mode is unstable. For $10B^2-1<0$ the first shape mode also becomes unstable. Moreover, for $12B^2-1<0$ the soliton becomes unstable against the interaction with radiation modes (Fig. \ref {Fig.12}). In Figure \ref{Fig.13} we present the case $4B^2>1$. Here the amplitude of radiation mode is greater than the case presented in Fig. \ref{Fig.12}. Notwithstanding, the original shape of the kink is recovered. The continuous spectrum (which concerns with radiation) was considered in our calculations and numerical experiments. In the numerical experiments there are emission of radiation that is absorbed by the resistance (dissipative terms). When the soliton is trapped the translational mode has a relevant role in contrast with the case for which the free soliton enters in interaction with an impurity. For the last case the shape modes indeed play a determinant mode. \section{Summary and Conclusions} In this paper we have presented the importance of considering the soliton as an extended particle as it interacts with real inhomogeneous media. We have shown that a finite-width impurity can activate a large number of soliton internal modes. In fact, for an impurity of the $N(x)\phi $ type with a stable point the number of mode increases (as the width of the impurity increases) in comparison with the free soliton case. Surprisingly, this contrasts with the inhomogeneous external force $F(x)$ with a stable point case, where the opposite occurs. An interesting finite-size effect occurs when both inhomogeneities, $F(x)$ and $N(x)\phi $, exhibiting only one equilibrium point when considered individually, generate a double-well effective potencial for the soliton as they interact between each other and with the soliton. In addition, we have predicted surprising effects when more than one impurity is considered. For instance, the stability condition depends on a variety of parameters that we can control (i.e., the width and the height of the impurity as well as the width of the soliton) and which we present as very relevant to the whole dynamics. We have shown the importance of the internal modes of the soliton as they can generate shape chaos for the soliton as well as cases in which the first shape mode leads the dynamics. Our approach has been also applied to the stochastic resonance\cite{g:51} of solitons that can oscillate in a double-well potential\cite{g:5}. Our approach provides the possibility to encounter resonances linked to the different internal modes of the soliton as well as to tune the time-dependent driving force to a selected mode. The results presented in this paper are very relevant for the concrete physical systems presented in Sec. II. as they concern the pinning problem (i.e., the stability problem of the soliton in an equilibrium point created by the inhomogeneities). As we have seen, it is not trivial to determine whether the equilibrium position is stable or not for the soliton when we are in the presence of external fields and inhomogeneities having finite widths and, also the extended character of soliton is considered. Real experiments\cite{g:6,g:7} where this kind of phenomena has been observed show that, in certain cases, the description of the impurities using delta functions can lead to erroneous conclusions. Soliton oscillators in Josephson junctions have been studied intensively in the last years due to their application as sources of radiation\cite{g:8}.\ In this context the study presented in this paper of the sustained oscillations of a soliton in an effective potential well has great importance. Gr\o nbech-Jensen and Blackburn\cite{g:9} have studied a system of coupled Josephson junctions as a super-radiant power source. Their point is that the oscillator ceases to be a point-like oscillator and exhibits an spatial degree of freedom. This leads to a radiation power greater than the expected from the point-like oscillators theory. In this paper we have shown that under the effects of the inhomogeneities the soliton exhibit an internal spatial degree of freedom. Furthermore, there is an increased number of activated internal modes. Particularly, we have shown that for impurities of the type $N(x)\phi $ with a stable equilibrium position, the hyperradiant effect could be achieved varying the impurity width. In this case the width of soliton will be greater than the width of the unperturbed soliton. This is also valid for impurities of the type $N(x)\sin \phi $ which can describe a microshort in a sine-Gordon-like systems. The chaotic oscillations and the soliton explosion phenomanon (due to soliton instability against the interaction with radiation modes) can be also important for the design of soliton based devices. \acknowledgments This work has been partially supported by Consejo Nacional de Investigaciones Cient\'ificas y Tecnol\'ogicas (CONICIT) under Project S1-2708.
\section{Observations and Data Reduction} Observations in J and Ks bands have been carried out at the ESO 3.5 m New Technology Telescope (NTT) with the new infrared imager/spectrometer SOFI (Moorwood et al. 1998). It is equipped with a 1024$\times$1024 pixel Rockwell Hawaii array (0.292 arcsec/pix). The observed field, centered at $\alpha=12^h:05^m:26.02^s$, $\delta=-07^o:43':26.43''$ (J2000) include the NTT Deep Field (Arnouts et al. 1999). The observations have been performed with seeing conditions ranging from 0.6-0.9 arcsec (median FWHM = 0.75 arcsec). A total exposure time of 624 min in Ks and 256 min in J was achieved as a result of the jittered frame co-addition. Raw data have been corrected for dark-current pattern by subtracting for each night a median dark and for pixel-to-pixel gain variations using a differential dome flat field. After flat fielding and standard sky subtraction, each basic set of frames has been registered and finally co-added to produce the final image The whole procedure has been accomplished relying on the original {\it Jitter} package by Devillard (1998; http://www.eso.org/eclipse). \begin{table*} \caption{ Summary of observations and image quality.} \begin{center} \begin{tabular}{cccccc} \hline \hline Filter & Number of & t$_{exp}$& $<FWHM>$& $\mu$&m$_{lim}$\\ & frames & (s) &(arcsec)&mag/arcsec$^2$& (3$\sigma$)\\ \hline Ks & 624 & 37440 & 0.75& 23.95&22.76\\ J & 128 & 15240 & 0.75& 25.75&24.57\\ \hline \hline \end{tabular} \end{center} \end{table*} The limiting surface brightness as well as the 3$\sigma$ magnitude within a 2$\times$ FWHM (1.5 arcsec) circular aperture estimated on the final images are reported in Tab. 1. \section{Object Detection and Magnitudes} The extraction and the magnitude estimate of sources have been performed with the SExtractor image analysis package (Bertin \& Arnouts, 1996). The object detection has been performed in the central deepest portion (4.45$\times$4.45 arcmin) of the images by first convolving them with a gaussian function having a FWHM=2.5 pixels ($\sim0.73$ arcsec). A 1$\sigma_{bkg}$ detection threshold and a minimum detection area equal to the seeing disk have been used. This rather low detection threshold provides a raw catalog from which we have extracted different samples by varying the S/N limit in order to optimize completeness and number of spurious detections for a given magnitude limit. For example, at a $S/N=5$ detection limit, compact objects as faint as $Ks = 23.0$ and $J = 24.8$ were singled out in our images, compared with $Ks = 23.9-24.5$ and $J=24.9-25.2$ in the deepest Keck observations (Bershady et al. 1998; Djorgovski et al. 1995). We are therefore confident that our sample is not biased against the detection of small compact sources, being fully comparable with the deepest data previously obtained by other authors on smaller areas on the sky (Gardner et al. 1993; Cowie et al. 1994; Soifer et al. 1994; McLeod et al. 1995; Djorgovski et al. 1995; Moustakas et al. 1997; Minezaki et al. 1998). We have measured the reliability of the detections by evaluating the number of spurious sources and their magnitude distribution using a small set of noise images obtained by combining the corresponding set of frames with ``wrong'' shifts. The images thus obtained do not contain sources and are characterized by the same mean background noise of the original data. We have deduced that a S/N$\ge$5 cutoff seems to ensure the optimum threshold for confident object detection both in J and Ks within a 0.3 mag accuracy. On the basis of the above selection criterium our subsequent analysis will rely on the {\it bona fide} Ks and J samples containing 1025 sources down to Ks$\le22.5$ and 1569 sources down to J$\le$24.0, respectively. Magnitudes have been estimated within a diameter aperture of 2.5 arcsec ($\sim3\times$FWHM) and corrected to ``total'' using an aperture correction of 0.25 mag obtained using a sample of bright objects. \section{Results} \subsection{Differential Galaxy Counts} In order to derive differential galaxy counts, we have first ``cleaned'' the two samples of stars. We have defined as stars those sources brighter than Ks$=19$ (in the Ks-band selected sample) and J$=20$ (in the J-band selected sample), having a value of SExtractor ``stellarity'' index larger than 0.9 both in the J sample and in the Ks one. \begin{figure*} \centerline{\psfig{figure=kcounts_bn.ps,height=70mm}} \caption{The figure compares the Ks-band counts obtained in this work with those in the literature: Djorgovski et al. (1995, Djo95), Bershady et al. (1998, Ber98) (obtained at the Keck telescope), Gardner et al. (1993, HMWS, HMDS, HDS), Glazebrook et al. (1994, Gla94), Soifer et al. (1994, Soi94), McLeod et al. (1995, McL95), Moustakas et al. (1997, Mou97), Saracco et al. (1997, ESOKS1, 2) and Minezaki et al. (1998, Min98) NICMOS/HST H-band counts of Yan et al. (1998, Yan98) are also shown.} \end{figure*} Then we have evaluated the completeness correction to apply to the raw data to account for the number of sources undetected at a given magnitude. This correction is obviously stronger at faint magnitudes and it mainly depends on the source spatial structure. Thus, we have generated a set of frames by dimming the final J and Ks frames themselves by various factors while keeping constant the background noise. This way we have used the whole real sample of sources reproducing the whole range of size and shape. SExtractor was then run with the same detection parameters to search for sources in each dimmed frame. Fig. 1 shows the Ks-band galaxy counts here derived superimposed to those in the literature. Our counts follow a d$logN/dK$ relation with a slope of 0.38 in the magnitude range $17<K<22.5$ and do not show any evidence of a turnover or of a flattening down to the limits of the survey. In Fig. 2 our J-band galaxy counts are compared with those of Bershady et al. (1998), the only J-band data available in the literature. We obtain a slope of J-band counts $\sim$0.36 in the magnitude interval $18<J<24$. Also the J-band counts do not show any hint of a decrease down to the limits of the data. \begin{figure} \centerline{\psfig{figure=jcounts_bn.ps,height=70mm}} \caption{The J-band galaxy counts obtained in this work are compared with those previously obtained by Bershady et al. (1998, Ber98): the counts continue to rise with a power law slope of $\sim$0.36.} \end{figure} The total numbers of galaxies represented in Fig.1 and 2 are 874 for the Ks-band and 1285 for the J-band and represent the largest samples at these depths. \subsection{Color and Size of Galaxies} The 2.5 arcsec aperture J-K color of galaxies in our sample was obtained by running SExtractor in the so-called {\em double-image mode}: J magnitudes for the Ks selected sample have been derived using the Ks frame as reference image for detection and the J image for measurements only, {\it vice versa} for the J selected sample. In Fig. 3 and 4 the J-Ks color of each galaxy is shown as a function of the apparent magnitude for the two samples. The median J-Ks color in each magnitude bin is also over-plotted as filled circles. The error bars are the standard deviation from the mean of the values within each bin. \begin{figure} \centerline{\psfig{figure=kcolor_can.ps,height=70mm}} \caption{J-Ks color of K-band selected galaxies as a function of Ks magnitude. The filled symbols represent the median J-Ks color of galaxies in each magnitude bin. Vertical error bars are the standard deviation from the mean of the values within each bin. The width of the bins are represented by the horizontal error bars The small crossed symbols, distributed along the right hand line, represent the sources undetected in the J frame, i.e. the J-Ks lower limits. } \end{figure} Both the samples show first a reddening trend of the median J-Ks color down to Ks$\sim19$ and J$\sim20$ respectively. Fainter than these magnitudes the J-K color of galaxies follows an unexpected behavior describing a bluing trend with increasing apparent magnitudes. \begin{figure} \centerline{\psfig{figure=jcolor_can.ps,height=70mm}} \caption{J-Ks color of J-band selected galaxies as a function of J magnitude. Symbols are as in Fig. 3. } \end{figure} In order to measure the size of galaxies of our samples, we have made use of the metric size function $\eta(\theta)$ introduced first by Petrosian (1976) \begin{equation} \eta(\theta)={{1~d~ln~l(\theta)}\over{2~d~ln(\theta)}} \end{equation} (Kron 1993), where $l(\theta)$ is the growth curve. This function has the property $\eta(\theta)={{I(\theta)}/{\langle I\rangle}_\theta}$ where $I(\theta)$ is the surface brightness at radius $\theta$ and ${\langle I\rangle}_\theta$ is the mean surface brightness within $\theta$. The function $\eta(\theta)$ has been obtained for each galaxy by constructing its own intensity profile. Following Bershady et al. (1998) we have defined the angular size of galaxies the value $\theta_{\eta}$ such that $\eta(\theta_{\eta})=0.5$. The median $\theta_{0.5}$ values are 0.73 arcsec and 0.79 arcsec for the J-band and Ks-band selected sample respectively. We have then measured the compactness of galaxies by defining the luminosity concentration index $C_\eta$ \begin{equation} C_\eta={{F(<\theta_{0.5})}\over{F(<1.5\theta_{0.5})}} \end{equation} that is the ratio between the flux within the radius $\theta_{0.5}$ and the flux interior to the radius $1.5\theta_{0.5}$. In Fig. 5 the concentration index $C_\eta$ and the apparent radius $\theta_{0.5}$ of galaxies belonging to the Ks sample are shown as a function of the apparent magnitude Ks in the upper and lower panel respectively. \begin{figure} \centerline{\psfig{figure=kdimen_bn.ps,height=70mm}} \caption{Light concentration index $C_\eta$ (upper panel) and apparent size $\theta_{0.5}$ (lower panel) of galaxies as a function of their Ks magnitude. The filled circles are the median values in 0.5 magnitude width bin. Vertical error bars are the standard deviation from the mean of the values within each bin while horizontal bars represent the width of the bins. The solid line in the lower panel shows the magnitude enclosed in an aperture of radius $\theta$ and a uniform surface brightness Ks=22.76 mag/arcsec$^2$. } \end{figure} The light concentration index seems to remain constant down to Ks$\sim19$ and J$\sim20$ while it is evident that galaxies become systematically more compact to fainter magnitudes. This trend is clearly present in both the samples as shown by the decrease of the median values $C_\eta$ with increasing apparent magnitudes. It is worth noting that the increasing compactness of galaxies occurs in the same magnitude ranges where galaxies get bluer. Moreover this trend seems to be matched also by a decrease of the apparent size of galaxies. \section{Conclusions} The main results obtained by this preliminary analysis of the data are the following: \begin{itemize} \item number counts follow a $d~log(N)/dm$ relation with a slope of 0.38 in Ks and of 0.36 in J showing no sign of flattening or turnover down to the faintest magnitudes. Such slopes and behavior fully agree with most of the ground-based data and with the deepest NICMOS/HST data (Yan et al. 1998); \item fainter than Ks$\sim19$ and J$\sim20$, the median J-K color of galaxies shows a break in its reddening trend turning toward bluer colors; \item faint bluer galaxies also display a larger compactness index, and a smaller apparent size. \end{itemize} The absence of a turnover in the counts down to the faintest magnitudes is indicative of an increasing contribution of sub-L$^*$ and hence local ($z \ll 1$) galaxies leading to a substantial steepening in the faint-end tail of the galaxy IR LF. Such a claim is also supported by the observed color trend which seems to rule out any major contribution of high-$z$ galaxies as primary contributors to faint counts. The existence of a steep ($\alpha\ll-1$) faint-end tail ($L<0.1-0.01L^*$) in the IR LF would naturally account both for the J-K color trend and the compactness trend shown by galaxies.
\section{Introduction} X-ray photoemission spectroscopy PES is a useful tool for studying the electronic structure of solids. The theoretical description of PES is however very complicated\cite{Hedinbook,Schaich} and almost all work has been based on the so-called sudden approximation.\cite{sudden,Lars98} The photoemission spectrum is then described by the electron spectral function convoluted by a loss function, describing the transport of the emitted electron to the surface. The sudden approximation becomes exact in the limit when the kinetic energy of the emitted electron becomes infinite.\cite {sudden,Lars98} In this limit we can distinguish between intrinsic satellites, appearing in the electron spectral function, and extrinsic satellites, appearing in the loss function. For lower kinetic energy, this distinction is blurred due to interference effects, and the satellite weights are expected to be quite different as we approach the opposite limit, the adiabatic limit, of low kinetic energy.\cite{sudden} It is then interesting to ask at what kinetic energy the sudden approximation becomes accurate. This issue has been studied extensively for the case when the emitted electron couples to plasmons, and it has been found that the sudden approximation becomes valid only for very large ($\sim $ keV) kinetic energies.\cite{Lars98,plasmon,plasmonexp} A semi-classical approach has been found to work exceedingly well for the study of plasmon satellites.\cite{Lars98,plasmon} In such a picture, one may take the emitted electron to move as a classical particle away from the region where the hole was created. The system then sees the potential from both the created electron and hole. Initially the electron potential cancels the hole potential, but as the electron moves away, the hole potential is gradually switched on. The switching-on of the hole potential may lead to the creation of excitations. If the kinetic energy of the electron is sufficiently large, we can consider the hole potential as being switched on instantly, and the creation of excitations around the hole then reaches a limiting value, the sudden limit. In the semi-classical picture, the perturbation is turned on during a time $\tau =R_{0}/v$, where $v$ is the photoelectron velocity and $R_{0}$ is the range of the interaction (scattering) potential between the emitted electron and the excitations. In this picture we also need to determine the relevant time scale $\tau _{{\rm max}}$ so that the sudden limit is reached if $\tau \ll \tau _{{\rm max}}$ or $v\gg R_{0}/\tau _{{\rm max}}$. Our analysis within the semi-classical framework shows that $1/\tau _{{\rm max}}$ is related to the energy $\delta E $ of the relevant excitation of the system and to the strength $\tilde{V}$ of the scattering potential. We find a different characteristic energy scale $\tilde E_s= 1/(2 \tilde R_s^2)$, where $\tilde R_s$ is a characteristic length scale of the scattering potential. On dimensional grounds one may argue that the adiabatic-sudden approximation takes place when the kinetic energy of the emitted electron is comparable to $\tilde E_s$. This would differ dramatically from the semi-classical approach, where the transition takes place for energy of the order $1/(\tilde E_s \tau_{{\rm max}}^2)$, i.e. e.g., $(\delta E)^2/\tilde E_s$ or $\tilde V^2/\tilde E_s$. Alternatively, and again on dimensional grounds, one may argue that the sudden approximation becomes valid when the kinetic energy of the emitted electron is much larger than the energy $\delta E$ of the relevant excitations of the system,\cite{Krause1974} in strong contrast to the two criteria above. This latter criterion is however not true in general.\cite{Shirley1987} For many systems with strong correlations, the core level spectrum can be understood in a charge transfer scenario.\cite{Kotani} This is illustrated in Fig. 1 for a Cu compound, e.g., a Cu halide. In the ground-state, Cu has essentially the configuration d$^{9}$ and all the ligand orbitals are filled. In the presence of a Cu core hole, it becomes energetically favorable to transfer an electron from a ligand to the d-shell, obtaining a $% d^{10}$ configuration on the Cu atom with the core hole. Due to the hybridization between the d$^{9}$ and $d^{10}$ configurations, the states are actually mixtures of the two configurations, as indicated in Fig. \ref {fig:schematic}. In the photoemission process there is a nonzero probability that the outer electron will not stay on the ligand, but is transferred to the lower energy $d$-like state. This ``shake-down'' process corresponds to the leading peak in the spectrum, while the process where the outer electron stays on the ligand corresponds to the satellite. This kind of model has been applied to rare earth compounds,\cite{Kotani,Gunnar83} chemisorption systems,\cite{Schonhammer,Fuggle} transition metal compounds,\cite {Larsson,Wallbank,Laan,Boer,Fujimori} and High $T_{c}$ compounds.\cite {Veenendaal} \begin{figure} \vspace*{4cm} \special{psfile=FIG1.ps} \caption{Schematic view of the Cu $3s$ charge transfer photoemission. Here $a $ is the Cu $3d$ level and $b$ is a ligand (L) valence state for the symmetric case. } \label{fig:schematic} \end{figure} Our simple model allows an accurate numerical calculation of the photocurrent either by integrating the time-dependent Schr\"{o}dinger equation or by directly inverting a resolvent operator (QM). We also derive analytic results with a separable potential. These results are compared with the semi-classical theory (SC), and with first order perturbation theory (PT). In both these cases we have analytic results, which is very useful for understanding the physics of the problem. The impurity model discussed here differs in certain important aspects from a real solid. To start with, for a solid we never reach the limit of a pure intrinsic spectrum since when the cross section for extrinsic scattering goes to zero, the range from which the photoelectrons come goes to infinity. In our impurity model, on the other hand, the extrinsic scattering approaches zero at high kinetic energies. Secondly, for a solid, we discuss excitations in the continuum and not as here discrete energy levels. For the coupling to plasmons, the adiabatic-sudden transition takes place at large kinetic energies where the SC approximation is very accurate.\cite{Lars98,plasmon} The relevant length scale is given by the plasmon wave length $\lambda =2\pi /q$ and the relevant time by the inverse plasmon frequency $\omega _{q}$. Large interference effects are then connected with a large phase velocity $\omega _{q}/q$, as discussed e.g. by Inglesfield.\cite{Inglesfield} Since long wave-length plasmons play an important role these large interference effects for small $q$ delay the approach to the sudden limit, which only is reached at very high kinetic energies ($\sim $ keV). For the localized excitations studied here the relevant length scale is much shorter. The SC theory then predicts that the transition takes place at correspondingly smaller kinetic energies. This is indeed what we find from the exact solution of our model. This has two consequences. Firstly, the SC treatment itself is not valid at such small energies, and we have to rely on QM treatments. Actually, although the SC treatment correctly predicts a small transition energy, we find it predicts qualitatively wrong dependencies on the relevant parameters. Secondly, the smaller energy scale means that the energy variation of the dipole matrix elements becomes very important. The dipole matrix element grows rapidly on an energy scale $% \tilde E_d$, which can become very important for the adiabatic-sudden transition. We study the ratio $r(\omega )$ between the weights of the satellite and the main peak as a function of the photon energy $\omega $ for the emission from a $3s$ level. First we consider the case when the scattering potential between the electron and the target is neglected. We find that the corresponding ratio $r_{0}(\omega )$ strongly depends on the ratio between the excitation energy $\delta E$ and $\tilde{E}_{d}$. If $\delta E/\tilde{E}_{d}\ll 1$, $r_{0}(\omega )$ approaches its limiting value from below, while it has an overshoot if $\delta E/\tilde{E}_{d}\gg 1$. In both case the limit value is reached for photoelectron energies of the order of a few times $\delta E$. We then study the effects of the scattering potential by focusing on $% r(\omega )/r_{0}(\omega )$. For small energies there is typically an overshoot due to {\it constructive} interference in the ''shake-down'' case, contrary to the shake-up case where, like for plasmons, $r(\omega )/r_{0}(\omega )$ is reduced by interaction effects. This happens on the energy scale $\tilde{E}_{s}$. If the scattering potential is very strong, this overshoot may extend to several times $\tilde{E}_{s}$. Depending on the parameters there may be an undershoot for higher energies, which can extend up to quite high energies. The undershoot is, however, rather small for the parameters we consider here, and should therefore not be very important unless we want to calculate the spectrum with a high accuracy. For Cu compounds and emission from the $3s$ core level, we find that $\tilde E_d$ and $\tilde E_s$ are comparable, and the relevant energy for $r(\omega)/r_0(\omega)$ is then given by $\tilde E\sim \tilde E_s\sim \tilde E_d$. We present our model in Sec. II and calculate various matrix elements in Sec. III. The sudden approximation is described in Sec. IV and exact numerical methods are given in Sec. V. The perturbational and the semi-classical treatments are presented in Sec. VI. In Sec. VII we study the condition for the adiabatic-sudden transition qualitatively, using simple analytic matrix elements and within the framework of the semi-classical theory. The results are discussed in Sec. VIII. \section{Model} We consider a Hamiltonian ${\cal H}_{0}$ describing a model with a core level $c$ and two valence levels $a$ and $b$, \begin{eqnarray}\label{eq:1} {\cal H}_{0} &=&\epsilon _{a}n_{a}+\epsilon _{b}n_{b}+\epsilon _{c}n_{c}+U_{a}n_{c}n_{a}+U_{b}n_{c}n_{b} \nonumber \\ &+&t(c_{a}^{\dagger }c_{b}+c_{b}^{\dagger }c_{a}). \end{eqnarray} The first two terms give the bare energies of the levels $a$ and $b$, and the last term the hybridization between them. The remaining terms involve the occupation number $n_{c}$ of the core level $c$. In photoemission the core level is filled in the initial state, and empty in the final, and $n_{c} $ only enters as a constant. It is trivial to diagonalize ${\cal H}_{0}$, and one obtains two dressed energies $E_{a}(n_{c})$ and $E_{b}(n_{c})$ for the levels $a$ and $b.$ In a Cu compound, for instance, $c$ may represent the Cu $3s$ core level, $a$ the Cu $3d$ valence level and $b$ a ligand state. This is schematically illustrated in Fig. \ref{fig:schematic}. In our calculations we almost always treat the case when $E_{a}(1)>E_{b}(1)$, and $% E_{b}(0)>E_{a}(0)$. The meaning of the levels $a$ and $b$ for different types of systems with localized excitations is indicated in Table \ref {general-CT}. The full Hamiltonian also has a one-electron part for continuum states, \begin{equation} T=\sum_{{\bf k}}\epsilon _{{\bf k}}n_{{\bf k}}, \label{eq:2} \end{equation} with the energies $\epsilon _{{\bf k}}=k^{2}/2$, and wavefunctions $\psi _{{\bf k}}$ obtained from a one-electron potential corresponding to $n_{c}=0$% . The $n_{{\bf k}}$ are occupation numbers $n_{{\bf k}}=c_{{\bf k}}^{\dagger }c_{{\bf k}}$. We use atomic units with $e=m=\hbar =1,$ and thus, e.g. energies are in hartrees ($27.2$ eV). The perturbation causing photoemission is \begin{equation} \Delta =\sum_{{\bf k}c}(M_{{\bf k}}c_{{\bf k}}^{\dagger }c_{c}+h.c.), \label{eq:3} \end{equation} where $M_{{\bf k}}$ is an optical transition matrix element. We take the photoelectron interaction as \begin{equation} V=\sum_{{\bf k}{\bf k}^{^{\prime }}}[n_{a}V_{{\bf k}{\bf k}^{^{\prime }}}^{(a)}+n_{b}V_{{\bf k}{\bf k}^{^{\prime }}}^{(b)}-V_{{\bf k}{\bf k}% ^{^{\prime }}}^{(c)}]c_{{\bf k}}^{\dagger }c_{{\bf k}^{^{\prime }}}. \label{eq:5} \end{equation} Here $V_{{\bf k}{\bf k}^{^{\prime }}}^{(\nu )}$ is a matrix element of the Coulomb potential $V^{(\nu )}({\bf r})$ from the charge density $\rho _{\nu }({\bf r})$ of the orbital $\nu $, \[ V_{{\bf k}{\bf k}^{^{\prime }}}^{(\nu )}=\int \psi _{{\bf k}}^{\ast }\left( {\bf r}\right) V^{(\nu )}({\bf r})\psi _{{\bf k}^{\prime }}\left( {\bf r}% \right) d{\bf r,\;}V^{(\nu )}({\bf r})=\int {\frac{\rho _{\nu }({\bf r}% ^{\prime })}{|{\bf r}-{\bf r}^{\prime }|}}d{\bf r}^{\prime }. \] It is the potential $V$ which determines the transition from the adiabatic to the sudden limit, with $V=0$ we are in the sudden limit. The total Hamiltonian is given by \begin{equation} {\cal H}={\cal H}_{0}+T+V+\Delta . \label{eq:6} \end{equation} This Hamiltonian has two conserved quantities, \begin{equation} n_{c}+\sum_{{\bf k}}n_{{\bf k}}=1\ \ \ {\rm and}\ \ \ n_{a}+n_{b}=1. \label{eq:7} \end{equation} For simplicity we take the core electron and the $d$ electron potentials as equal, $V^{c}=V^{a}$, and use the relation $n_{a}+n_{b}=1$ to obtain \begin{equation} V=n_{b}\sum_{{\bf k}{\bf k}^{^{\prime }}}V_{{\bf k}{\bf k}^{^{\prime }}}c_{ {\bf k}}^{\dagger }c_{{\bf k}^{^{\prime }}} \label{eq:8} \end{equation} where \begin{eqnarray*} V_{{\bf kk}^{^{\prime }}} &\equiv &\int \psi _{{\bf k}}^{\ast }\left( {\bf r} \right) V_{sc}({\bf r})\psi _{{\bf k}^{\prime }} \left( {\bf r}\right) d{\bf r,\;} \\ V_{sc}({\bf r}) &\equiv &V^{(b)}({\bf r})-V^{(a)}({\bf r})=\int d{\bf r} ^{\prime }{\frac{1}{|{\bf r}-{\bf r}^{\prime }|}}[\rho _{b}({\bf r}^{\prime })-\rho _{a}({\bf r}^{\prime })]. \end{eqnarray*} The (scattering) potential $V_{sc}({\bf r})$ describes the change in the potential acting on the emitted electron when the electron in the target hops from level $a$ to $b$. Dropping a constant we can write ${\cal H}_{0}$ as \begin{eqnarray}\label{eq:9} {\cal H}_{0} &=&\epsilon _{c}n_{c}+(\epsilon _{a}+Un_{c})n_{a}+\epsilon _{b}n_{b} \nonumber \\ &+&t(c_{a}^{\dagger }c_{b}+c_{b}^{\dagger }c_{a}). \end{eqnarray} where the Coulomb integral $U$ is given by \begin{eqnarray}\label{eq:m1} U &\equiv &U_{a}-U_{b} \nonumber \\ &=&\int d{\bf r}d{\bf r}^{\prime }\rho _{c}({\bf r}){\frac{1}{|{\bf r}-{\bf r% }^{\prime }|}}[\rho _{a}({\bf r}^{\prime })-\rho _{b}({\bf r}^{\prime })]. \end{eqnarray} Since the core level is very localized in space this leads to \begin{equation} U=-V_{sc}(0). \label{eq:m3} \end{equation} For the different types of systems in Table \ref{general-CT}, $a$ refers to a localized level and $b$ refers to a more extended level. For instance, for a copper dihalide compound, $a$ refers to a Cu $3d$ orbital and $b$ to a combination of orbitals on the ligand sites. For simplicity, we approximate the six ligand orbitals by a spherical shell with the radius $R_{0}$,\cite {Kaupp} where $R_{0}$ is the average Cu-ligand separation. The potential from the Cu $3d$ orbital $V_{3d}(r)$ can be considered as purely Coulombic at $r=R_{0}$. The charge from the spherical shell gives a constant potential inside the radius $R_{0}$, and we have \begin{equation} \label{eq:m4} V_{sc}(r)=\left\{ \begin{array}{ll} (-V_{3d}(r)+\frac{1}{R_{0}})/\varepsilon & r<R_{0}; \\ 0 & r>R_{0}. \end{array} \right. \end{equation} Here $\varepsilon $ is a constant chosen to make $U=-V_{sc}(0)$, which may be thought of as being due to screening by the surrounding. Since $V_{3d}(0)\gg 1/R_{0}$, $\varepsilon $ varies only weakly with $R_{0}$ and is approximately given by $\varepsilon \simeq V_{3d}(0)/U$. \section{Matrix elements} \label{sec:matrix} To estimate the matrix elements $M_{{\bf k}}$ and $V_{{\bf k}{\bf k}% ^{^{\prime }}}$ we must approximate the photoelectron wavefunctions $\psi _{{\bf k}}\left( {\bf r}\right) $. These wavefunctions are calculated from the potential of a neutral atom, which further is shifted to make the potential zero outside a muffin-tin radius $r_{mt}$. The states are then described by spherical Bessel functions outside $r_{mt}$, which are matched to a solution of the atomic potential inside $r_{mt}$. For the energy $k^{2}/2$ we obtain the partial wave \begin{equation} R_{lk}(r)=\cases {a_{lk}\psi_{lk}(r) & $r<r_{mt}$;\cr \sqrt{\frac{2}{R}}k[\cos\eta_{lk}j_l(kr)-\sin\eta_{lk}n_l(kr)] & $r>r_{mt}$,\cr} \label{eq:ma1} \end{equation} where $\psi _{lk}(r)$ is the solution of the radial Schr\"{o}dinger equation for the atomic potential inside the muffin-tin radius, $a_{lk}$ is a matching coefficient and $\eta _{lk}$ a phase shift. The normalization is given by \begin{equation} \int_{0}^{R}drr^{2}R_{lk}^{2}(r)=1, \label{eq:ma1a} \end{equation} where $R$ is the radius of a large sphere to which the continuum states are normalized. The factor $(2/R)^{1/2}k$ is due to the normalization and the asymptotic behavior of $j_{l}(x)$ for large $x$. Slater's rules\cite{Slater} are used to generate the orbitals and charge densities, from which the potential $V_{3d}(r)$ is calculated. This gives the scattering potential $V_{sc}(r)$, which is shown in Fig. \ref {fig:potential}a. We consider photoemission from a Cu $3s$ hole. Due to the dipole selection rules, the core electron is then emitted into a continuum state of $p$-symmetry. The matrix elements $V_{{k}{k}^{^{\prime }}}$ of the scattering potential $V_{sc}(r)$ are shown in Fig. \ref{fig:potential}b, \begin{equation} V_{kk^{^{\prime }}}=\int drr^{2}R_{k}(r)V_{sc}(r)R_{k^{^{\prime }}}(r), \label{eq:ma0} \end{equation} where the muffin-tin radius $r_{mt}$ is taken as the ionic radius of Cu, $% r_{mt}=2.6$ a.u., and we have dropped the $l$ index, since we always consider $l=1.$ The dipole matrix element $M_{k}$ is given by \begin{equation} M_{k}\sim a_{k}(\epsilon _{k}-\epsilon _{c})\int drr^{2}\psi _{3s}(r)r\psi _{k}(r). \label{eq:ma1b} \end{equation} We assume that the core level is deep, and $\left| \epsilon_{c}\right|$ much larger than the energy difference between the ligand and copper levels. We can then take the factor $\epsilon_{k}-\epsilon_{c}$ as a constant, which drops out since we always consider relative intensities. The result for $M_{k}$ is shown as the solid line in Fig. \ref{fig:dipole}a. These dipole and scattering potential matrix elements are used in the following numerical calculations. Extensive calculations of dipole matrix elements for many systems were performed by Yeh and Lindau.\cite{crossection} To interpret the results, it is useful to also perform analytical calculations. For this purpose we need models of the matrix elements. Below we consider the limits of low and high kinetic energies of the emitted electron. In the limit of low kinetic energies, we replace the spherical Bessel function by its expansion for small arguments \begin{equation} \label{eq:ma2} j_l(x)=\frac{1}{(2l+1)!!}x^l,\mbox{ }\mbox{ } n_l(x)=-(2l-1)!! \frac{1}{ x^{l+1}} \end{equation} and the solution $\psi_{lk}(r)$ by its zero energy limit $\psi_{l0}(r)$. This leads to \begin{eqnarray} \label{eq:ma3} \tan\eta_{lk}=\frac{l-\xi}{l+1+\xi}\frac{(kr_{mt})^{2l+1}} {(2l-1)!!(2l+1)!!} \sim\eta_{lk}, \nonumber \\ a_{lk}=\sqrt{\frac{2}{R}}k\left[\frac{2l+1}{l+1+\xi} \frac{(kr_{mt})^l}{ (2l+1)!!}\right]\frac{1}{\psi_{l0}(r_{mt})}. \end{eqnarray} \begin{figure} \vspace*{10.5cm} \special{psfile=FIG2.ps} \caption{(a) The photoelectron scattering potential $V_{sc}(r)$ given by Eq. (\ref{eq:m4}) with respect to $r$ for CuCl$_2$ ($R_0=4.71$ a.u. and $\protect% \varepsilon=1.96$). In the inset, we give the atomic configuration of Cu-Cl octahedral (nearly octahedral) cluster in CuCl$_2$. (b) The diagonal and off-diagonal matrix elements of the scattering potential multiplied by $R$. In the figure, $k_1=1$ a.u., $k_2=5$ a.u., and $k_3=10$ a.u. are taken. } \label{fig:potential} \end{figure} Due to the matching, the coefficient $a_{lk}$ contains the ratio $% \xi=r_{mt}\psi^{^{\prime}}_{l0}(r_{mt})/\psi_{l0}(r_{mt})$. The value of the coefficient therefore depends in an interesting way on the wave function $% \psi_{l0}$ and its derivative, and if $l+1+\xi$ is close to zero $a_{lk}$ blows up. Then the matrix elements of the scattering potential also blow up and we may expect strong deviations from the sudden approximation. In such a case the dipole matrix element $M_{lk}$ also becomes large, i.e., there is a resonance ($\eta_{lk}=\frac{\pi}{2}$) in the photoemission cross section. From Eqs.(\ref{eq:ma1}) and (\ref{eq:ma1b}) it follows that in the limit of a small $k$ the dipole matrix element is proportional to $a_{k}\sim k^{l+1}$% . The main peak and the satellites in the photoemission spectrum correspond to different kinetic energies and therefore have dipole matrix elements with different $k$-values. This is important at low energies, while for large energies the variation in $M_{lk}$ is generally small over a range of the energy difference between the main peak and the satellite, and for the ratio of the peaks, the dipole matrix elements should then not play a role. For simplicity, we assume that the dipole matrix elements become independent of $% k$ for $\tilde{R}_{d}k\gg 1$, where $\tilde{R}_{d}$ is some typical length scale of the system. For our case ($l=1$) we use the model (note that any constant factor in $M_{k}$ drops out in our final expressions) \begin{equation} M_{k}={\frac{(\tilde{R}_{d}k)^{2}}{1+(\tilde{R}_{d}k)^{2}}}\equiv {\frac{% \epsilon _{k}/\tilde{E}_{d}}{1+\epsilon _{k}/\tilde{E}_{d}}}, \label{eq:ma4} \end{equation} where $\tilde{E}_{d}=1/(2\tilde{R}_{d}^{2})$. Fig. \ref{fig:dipole}a compares this model with the full calculation for a $3s$ orbital. We obtain $% \tilde{R}_{d}=1.3$ a.u.. For a $1s$ or $2s$ orbital the length scale is smaller and $\tilde{R}_{d}\sim 1/2$ a.u.. While we consider $l=1$ the behaviour for other $l$ is primarily modified for small energies. We next consider the matrix elements $V_{kk^{^{\prime}}}$. For small values of $k$ and $k^{^{\prime}}$ and for $l=1$ \begin{eqnarray} \label{eq:ma5} V_{kk^{^{\prime}}}&=&\frac{2}{R}\left[\frac{r_{mt}}{\psi_{l0}(r_{mt})}\right] ^2 \frac{(kk^{^{\prime}})^2}{(\xi+2)^2}\int_0^{r_{mt}}\psi_{l0}^2(r)V_{sc} (r) r^2dr \nonumber \\ &+&\frac{2}{9R}(kk^{^{\prime}})^2\int_{r_{mt}}^{R_0} \left[r+\left(\frac{ 1-\xi}{2+\xi}\right)\frac{r_{mt}^3}{r^2}\right]^2 V_{sc}(r)r^2dr. \nonumber \\ \end{eqnarray} For small values of $k$ and $k^{^{\prime}}$ it then follows that $% V_{kk^{^{\prime}}}\sim (kk^{^{\prime}})^2$. For large values of $k$ and $% k^{^{\prime}}$ the matrix elements become very small due to destructive interference between the two wave-functions unless $k\approx k^{^{\prime}}$. If $k=k^{^{\prime}}$ the matrix elements $V_{kk}$ approach a constant. These features are contained in the model \begin{equation} \label{eq:ma6} V_{kk^{^{\prime}}}={\frac{\tilde V \tilde R_s}{R}}{\frac{ (\tilde R% _s^2kk^{^{\prime}})^2 }{ \lbrack 1+(\tilde R_s k)^2\rbrack \lbrack 1+(\tilde % R_s k^{^{\prime}})^2\rbrack \lbrack 1+\tilde R_{sd}^2(k-k^{^{\prime}})^2% \rbrack}}, \end{equation} where $\tilde R_s$ and $\tilde R_{sd}$ are appropriate length scales and $% \tilde V$ has the energy dimension. We recall that $\tilde V$ contains information about the coefficient $a_{lk}$ defined in Eq.(\ref{eq:ma3}) and therefore about the atomic potential and that $\tilde V$ may become particularly large close to a resonance. The dipole matrix element $M_k$ and potential matrix $V_{kk^{^{\prime}}}$ in our simplified model Eqs. (\ref {eq:ma4}) and (\ref{eq:ma6}) are given in Fig. \ref{fig:dipole} as compared with the exact results. For large values of $k$ and $k^{^{\prime }}$, the expression Eq.~(\ref {eq:ma6}) simplifies to \begin{equation} V_{kk^{^{\prime }}}={\frac{\tilde{V}\tilde{R}_{s}}{R}}{\frac{1}{1+\tilde{R}% _{sd}^{2}(k-k^{^{\prime }})^{2}}}. \label{eq:ma6c} \end{equation} An expression of this type can also be derived by assuming that the wave functions $R_{lk}(r)$ can be approximated by spherical Bessel functions in all of space, and by assuming some shape of $V_{sc}(r)$, e.g., a linear dependence on $r$ \begin{figure} \vspace*{10.5cm} \special{psfile=FIG3.ps} \caption{(a) The dipole matrix element $M_{k}$ as a function of $k$. The exact result (solid line) is obtained from Eq. (\ref{eq:ma1b}) and the simplified model by Eq. (\ref{eq:ma4}) is also shown (dashed line). The appropriate parameter is $\tilde R_d=1.30$ a.u. (b) The matrix elements $% V_{kk^{^{\prime}}}$ of the scattering potential are given for $% k=k^{^{\prime}}$, $k^{^{\prime}}=1$ a.u., and $k^{^{\prime}}=5$ a.u.. The solid line is from the exact calculation for the model of CuCl$_2$ and the dashed line is based on the simplified model (Eq.(\ref{eq:ma6})). The parameters are $\tilde{V}=-0.36$ a.u., $\tilde R_s=1.77$ a.u. and $\tilde R _{sd}=1.31$ a.u.. } \label{fig:dipole} \end{figure} \begin{equation} V_{sc}(r)=V_{sc}(0)(1-{\frac{r}{R_{0}}}). \label{eq:ma6a} \end{equation} For large values of $k$ and $k^{^{\prime }}$ we then obtain \begin{equation} V_{kk^{^{\prime }}}={\frac{V_{sc}(0)R_{0}}{R}}{\frac{1-{\rm cos}% [(k-k^{^{\prime }})R_{0}]}{[(k-k^{^{\prime }})R_{0}]^{2}}}. \label{eq:ma6aa} \end{equation} For this model we relate $\tilde{R}_{s}=\tilde{R}_{sd}=R_{0}/3$ to the range $R_{0}$ of the potential and $\tilde{V}=3V_{sc}(0)/2$. Using this identification in Eq. (\ref{eq:ma6c}), leads to the correct average value of $V_{kk}$ and to the correct width in $k-k^{^{\prime }}$ of $V_{kk^{^{\prime }}}$. The simple form (\ref{eq:ma6c}), however, neglects the effects of the oscillations of the cos-function in Eq. (\ref{eq:ma6aa}) for large values of $(k-k^{^{\prime }})R_{0}$, and it therefore gives a worse representation of the linear potential (\ref{eq:ma6a}) than the form (\ref{eq:ma6}) gives for the more realistic scattering potential (\ref{eq:m4}). Later we will find that it is a reasonable approximation to put $\tilde{R}_{d}$, $\tilde{R}_{sd} $ and $\tilde{R}_{s}$ equal to the same value $\tilde{R}$, and introduce the corresponding energy $\tilde{E}=1/(2\tilde{R}^{2})$. \section{Sudden approximation} \label{sec:sudden} We first discuss the photoemission in the sudden limit, i.e., we neglect the scattering potential between the emitted electron and the target ($V\equiv 0$). The initial state $|\Psi _{0}\rangle$ is the ground state of ${\cal H}_{0} $ with $n_{c}=1$ and given by \begin{equation} \label{eq:s1} |\Psi_0\rangle=-\sin\theta |\psi_c\rangle|\psi_a\rangle +\cos\theta |\psi_c\rangle|\psi_b\rangle, \end{equation} where \begin{equation} \tan 2\theta =2t/(\epsilon_{a}+U-\epsilon_{b}) \label{eq:s2} \end{equation} and the corresponding ground state energy is \begin{equation} \label{eq:s3} E_0=\epsilon_c+{\frac{1}{2}}(\epsilon_a+U+\epsilon_b)- \frac{1}{2}\sqrt{% (\epsilon _{a}+U-\epsilon _{b})^{2}+4t^{2}}. \end{equation} The final states of the target are given by the two eigenstates of ${\cal H} _0$ with $n_c=0$ \begin{eqnarray} \label{eq:s4} |\psi _{1}\rangle &=&\cos \varphi |\psi _{a}\rangle -\sin \varphi |\psi _{b}\rangle , \nonumber \\ |\psi _{2}\rangle &=&\sin \varphi |\psi _{a}\rangle +\cos \varphi |\psi _{b}\rangle , \end{eqnarray} with \begin{equation} \tan 2\varphi =2t/(\epsilon _{b}-\epsilon _{a}) \label{eq:s5} \end{equation} and the corresponding energy eigenvalues $E_{1}$ and $E_{2}$ are \begin{equation} E_{ {1 \atop 2} }={\frac{1}{2}}(\epsilon _{a}+\epsilon _{b})\mp \delta E/2, \label{eq:s6} \\ \end{equation} with \begin{equation} \delta E=\sqrt{(\epsilon _{a}-\epsilon _{b})^{2}+4t^{2}} \label{eq:s7} \end{equation} being the optical excitation energy of the system. The photocurrent $J_{k}^{s}(\omega )$ ($s=1,2$) is given by,\cite{Hedinbook} \begin{equation} J_{k}^{s}(\omega )=|\langle \Psi _{f}^{sk}|\Delta |\Psi _{0}\rangle |^{2}\delta (\omega -\epsilon _{k}+E_{0}-E_{s}), \label{eq:s8} \end{equation} where $|\Psi _{f}^{sk}\rangle $ is a final state. According to the sudden approximation, it can be written as the final target state multiplied by the photoelectron state, $|\Psi _{f}^{sk}\rangle =|\psi _{s}\rangle |\psi _{k}\rangle $. This gives \begin{equation} \langle \Psi _{f}^{sk}|\Delta |\Psi _{0}\rangle =M_{k}w_{s}\equiv m_{sk},\mbox{ }\mbox{ }w_{s}=\left\{ \begin{array}{c} -\sin \left( \varphi +\theta \right) ,\;s=1 \\ \cos \left( \varphi +\theta \right) ,\;\;s=2 \end{array} \right. \label{eq:s9} \end{equation} $J_{{\bf k}}^{1}(\omega )$ gives the main line (corresponding to the quasi particle line in metal) and $J_{{\bf k}}^{2}(\omega )$ the satellite line. The schematic picture of the initial and final state for this system is given in Fig. \ref{fig:schematic}. Summing the kinetic energy distribution of the photoelectron, we obtain the absorption spectra $J_{s}(\omega )$, \begin{eqnarray} \label{eq:s10} J_{s}(\omega ) &=&\sum_{k}J_{k}^{s}(\omega )\propto \frac{1}{k_{s}}% |M_{k_{s}}w_{s}|^{2} \nonumber \\ k_{s} &=&\sqrt{2(\omega +E_{0}-E_{s})}, \end{eqnarray} where the threshold energies for $J_{1}(\omega )$ and $J_{2}(\omega )$ are given by $E_{1}-E_{0}$ and $E_{2}-E_{0}(\equiv \omega _{th})$, respectively. We can thus also write $k_{1}=\sqrt{2(\omega +\delta E-\omega _{th})}$ and $% k_{2}=\sqrt{2(\omega -\omega _{th})}$. The factor $1/k$ comes from the $k$% -summation over a $\delta $-function in energy. For convenience we introduce the quantity $\tilde{\omega}=\omega -\omega _{th}$, and thus $\epsilon _{k_{2}}=\tilde{\omega}$. In the sudden approximation the kinetic energy of the emitted electron is large, and we can take $k_{1}=k_{2}$. The ratio $r_{00}$ of the satellite to the main peak intensity then is \begin{equation} r_{00}=\lim_{\omega \rightarrow \infty }\frac{J_{2}(\omega )}{J_{1}(\omega )}% =\cot ^{2}(\varphi +\theta ). \label{eq:s11} \end{equation} Taking into account the energy dependence of the dipole matrix element according to model Eq. (\ref{eq:ma4}) as well as the factor $1/k$, we obtain \begin{eqnarray}\label{eq:s12} r_{0}(\omega ) &=&r_{00}\left[ {\frac{\tilde{\omega}}{\tilde{\omega}+\delta E% }}\right] ^{\frac{3}{2}} \nonumber \\ &&\times \left[ {\frac{1+(\tilde{\omega}+\delta E)/\tilde{E}_{d}}{1+\tilde{% \omega}/\tilde{E}_{d}}}\right] ^{2}\Theta (\tilde{\omega}). \end{eqnarray} We now require that the ratio $r_{0}(\omega )$ should reach a fraction $% \gamma $ ($\gamma \approx 1$) of its limiting value $r_{0}(\infty )$ for $% \omega =\omega _{\gamma }$. This gives \begin{equation} {\frac{\omega _{\gamma }-\omega _{th}}{\delta E}}\approx \cases{{3\over 2}{1\over 1-\gamma}, & if $\delta E \ll \tilde{E}_d$;\cr {\gamma^{2/ 3} (\tilde{E}_d/\delta E)^{4/3}}, & if $\delta E \gg \tilde{E}_d$.\cr} \label{eq:s13} \end{equation} This criterion refers to the energy where $r_{0}(\omega )$ reaches a fraction $\gamma $ in its rising part, and it does not consider that there is a large overshoot for $\delta E/\tilde{E}_{d}\gg 1$. In this case we can instead require that $r_{0}(\omega )$ is smaller than $\gamma \approx 1$ in its descending part. This gives the condition \begin{equation} {\frac{\omega _{\gamma }-\omega _{th}}{\delta E}}\approx {\frac{1}{\gamma ^{2}-1}}\hskip0.5cm\delta E/\tilde{E}_{d}\gg 1,\gamma >1. \label{eq:s13a} \end{equation} In Fig. \ref{fig:sudden} we show results for $r_{0}(\omega )$ over a large range of values for $\delta E/\tilde{E}$. The figure illustrates that the dipole matrix element effect alone makes the sudden approximation invalid for small kinetic energies. It is interesting that for somewhat larger photon energies $r_{0}$ overshoots. The reason is that the matrix elements $M_{k}$ saturate for $\epsilon _{k}\gg \tilde{E}_{d}$, while the factor $1/k$ in Eq.(\ref{eq:s10}) favors the satellite. For $\delta E/\tilde{ E}_{d}=1$, the result is rather close to the sudden limit for $\tilde{\omega} /\delta E\sim 1$. Finally, for $\delta E/\tilde{E}_{d}\gg 1$, there is a substantial overshoot. \begin{figure}[bt] \unitlength1cm \begin{minipage}[t]{3.125in} \centerline{\rotatebox{-90}{\epsfxsize=2.5in \epsffile{FIG4.ps}}} \caption[]{\label{fig:sudden}The ratio $r_0(\omega)$ of the satellite to the main peak in Eq.(\ref{eq:s12}) divided by the result for an infinite photon energy ($r_0(\infty)=r_{00}$). Three values of the excitation energy $\delta E$ are considered.} \end{minipage} \hfill \end{figure} As discussed in the introduction, we would like to study how the adiabatic to sudden transition depends on certain factors, like the range $R_{0}$ of the potential and the energy $\delta E$ of the excitation causing the satellite. We therefore keep ratio of $t$, $U$ and $\epsilon _{a}-\epsilon _{b}$ fixed, but vary their magnitude. In this way we can vary $\delta E$ without varying the magnitude of the satellite in the sudden limit. Eq. (\ref {eq:m3}) requires that we vary $V_{sc}(0)$ as we vary $\delta E$ (via $U$), e.g., by varying the dielectric constant $\varepsilon $. In some of the calculations below, however, we do not impose Eq. (\ref{eq:m3}), to be able to see the effect of varying $\delta E$ alone. We furthermore vary the range $R_{0}$ of the potential. From the definition Eq. (\ref{eq:m4}) it follows that this would also vary the strength of the potential. For this reason we simultaneously vary the dielectric constant $\varepsilon $ so that $V_{sc}(0) $ stays unchanged when $R_{0}$ is changed. Alternatively, we can use the analytical matrix elements (\ref{eq:ma4}, \ref{eq:ma6}). We can then easily vary the length scale by changing $\tilde{R}$ or the strength by changing $% \tilde{V}$. To know roughly what are interesting values for our parameters we use experimental results for some copper dihalides.\cite{Laan} We estimate the relative strength of the satellite to the main peak, and the energy difference between the peaks. This gives two equations while in our model these quantities, $r_{00}$ and $\delta E$, depend on three parameters, $t,U,$ and $\epsilon _{a}-\epsilon _{b}$. To only have two parameters we consider {\it the symmetric case} $\epsilon _{a}=\epsilon -U/2$ and $\epsilon _{b}=\epsilon $ as shown in Fig.\ref{fig:schematic}. In the symmetric case we are restricted to the shake-down situation since before the transition the $a$-level is above the $b$-level, $% \epsilon _{a}+U-\epsilon _{b}=U/2$, while after the transition the $a$-level is below the $b$-level, $\epsilon _{b}-\epsilon _{a}=U/2$. In the symmetric case we have $0<\theta =\varphi <\pi /4$, and $r_{00}=\cot ^{2}2\varphi =U^{2}/(16t^{2})$. Once we know where in the ball-park we have $t$ and $U$, we can leave the symmetric case, and also consider, e.g., shake-up cases when there is no level crossing, $\epsilon _{a}>\epsilon _{b}$. In the lowest final state the electron essentially stays on level $b$, while the transfer of the electron to the level $a$ corresponds to a shake up satellite. In this case we have $-\pi /4<\varphi <0<\theta <\pi /4$ and $% \varphi +\theta <0$. Our calculations usually take the CuCl$_{2}$ parameters as reference values. For CuCl$_{2}$ we have $\theta =\varphi = 0.3$, which gives $r_{00}=2.1$. Further $\tilde{V}=-0.36$ a.u., $\tilde{E}=0.195$ a.u. (with $ \tilde{R}=1.6$ a.u.), and $\delta E=0.237$ a.u., i.e., $\tilde{V}/\tilde{E}=-1.85$ and $\delta E/\tilde{V}=-0.66$ (see also Table II). \section{Exact treatment} \subsection{Time-dependent formulation} To obtain exact results for model Eq.(\ref{eq:6}), we use a time-dependent formulation\cite{time-formalism} and solve the Schr\"odinger equation for the Hamiltonian \begin{equation} \label{eq:6a} {\cal H}(\tau)={\cal H}_0+T+V+\Delta f(\tau). \end{equation} The interaction is switched on at $\tau=0$, using \begin{equation} \label{eq:4} f(\tau)=e^{-i\omega \tau}(e^{-\eta \tau}-1) \hskip1cm \eta>0. \end{equation} Here $\eta$ is a small quantity to assure that the external field is switched on smoothly. The initial ($\tau=0$) state $|\Psi_0\rangle$ is given by the ground state of ${\cal H}_0$ with $n_c=1$ in Eq.(\ref{eq:1}). After a time $\tau$, the state $|\Psi(\tau)\rangle$ of the system is \begin{eqnarray} |\Psi(\tau)\rangle&=&a(\tau)|\psi_a\rangle|\psi_c\rangle +b(\tau)|\psi_b\rangle|\psi_c\rangle \nonumber \\ &+&\sum_k c_{ak}(\tau)|\psi_a\rangle|\psi_k\rangle +\sum_k c_{bk}(\tau)|\psi_b\rangle|\psi_k\rangle. \end{eqnarray} The coefficients of $|\Psi(\tau)\rangle$ can be determined by \begin{equation} i\frac{\partial}{\partial \tau}|\Psi(\tau)\rangle={\cal H} (\tau)|\Psi(\tau)\rangle, \end{equation} which gives four differential equations for the four coefficients $a(\tau)$, $b(\tau)$, $c_{ak}(\tau)$, and $c_{bk}(\tau)$, \begin{eqnarray} i\frac{\partial}{\partial \tau}a(\tau) &=&(\epsilon_a+U+\epsilon_c)a(\tau)+tb(\tau) \nonumber \\ &+&\sum_k{V_k^d}^{\ast}(\tau)c_{ak}(\tau), \end{eqnarray} \begin{eqnarray} i\frac{\partial}{\partial \tau}b(\tau) &=&(\epsilon_b+\epsilon_c)b(\tau)+ta(\tau) \nonumber \\ &+&\sum_k{V_k^d}^{\ast}(\tau)c_{bk}(\tau), \end{eqnarray} \begin{eqnarray} i\frac{\partial}{\partial \tau}c_{ak}(\tau) &=&(\epsilon_a+\epsilon_k)c_{ak}(\tau)+tc_{bk}(\tau) \nonumber \\ &+&V_k^d(\tau)a(\tau), \end{eqnarray} \begin{eqnarray} i\frac{\partial}{\partial \tau}c_{bk}(\tau) &=&(\epsilon_b+\epsilon_k)c_{bk}(\tau) +tc_{ak}(\tau)+V_k^d(\tau)b(\tau) \nonumber \\ &+&\sum_{k^{\prime}}V_{kk^{\prime}}c_{bk^{\prime}}(\tau), \end{eqnarray} where $V_k^d(\tau)=V_0M_kf(\tau)$ with $V_0$ representing the strength of the external field. We solve the equations in the limit when $V_0 \to 0$, and thus the ratio between $c_{ak}$ and $c_{bk}$ is independent of $V_0$. The initial conditions are $a(0)=-\sin\theta$, $b(0)=\cos\theta$, and $% c_{ak}(0)=c_{bk}(0)=0$. Thus the problem is reduced to solving the coupled differential equations, which is done using the Runge-Kutta fourth-order method. The photoelectron currents $J_1(\omega)$ and $J_2(\omega)$ corresponding to main and satellite lines, respectively, are given by \begin{eqnarray} \label{eq:t1} J_1(\omega)&=&\sum_k|\langle\Psi_f^{1k}|\Psi(\tau)\rangle|^2 \nonumber \\ &=&\sum_k|\cos\varphi c_{ak}(\tau)-\sin\varphi c_{bk}(\tau)|^2, \end{eqnarray} \begin{eqnarray} \label{eq:t2} J_2(\omega)&=&\sum_k|\langle\Psi_f^{2k}|\Psi(\tau)\rangle|^2 \nonumber \\ &=&\sum_k|\sin\varphi c_{ak}(\tau)+\cos\varphi c_{bk}(\tau)|^2, \end{eqnarray} where $\tau$ is a sufficiently large time. We let the system evolve for a time of the order $1/\eta$ to obtain converged results for a given finite $% V_0$. In principle, we should perform the calculation for a few small values of $\eta$ and then extrapolate to $\eta=0$ followed by an extrapolation $% V_0\to 0$. Here, for simplicity we have performed the calculation for one single small value of $V_0$. The calculation was performed for $\eta=0.1$ eV, 0.08 eV, 0.02 eV, and the results were extrapolated to $\eta=0$ assuming the $\eta$ dependence $a(\omega)\eta+b(\omega) \eta^2+c(\omega)$. The error in this approach occurs primarily for small $\tilde{\omega}(\lesssim 5$ eV), and it is then less than 5\% in $r(\omega)/r_0(\omega)$. The approach above gives the relative intensity of the main and satellite peaks \begin{equation} r(\omega)=\frac{J_2(\omega)}{J_1(\omega)}. \end{equation} It can be shown that the formulas (\ref{eq:t1}, \ref{eq:t2}) above give identical results to the more conventional formulation (\ref{eq:pt3}) below, by performing derivations of the type made in, e.g., Ref. \onlinecite{auger}. As an example of the results obtained in this formalism, we show in Fig. \ref {fig:Cu-dihalides} results for the copper dihalides CuBr$_{2}$, CuCl$_{2}$ and CuF$_{2}$. The corresponding parameters are shown in Table \ref {used-parameter} and were estimated from experiment.\cite{Laan} The figure illustrates that there is a small ``overshoot'' for small $\tilde{\omega}$ but that the sudden limit is reached fairly quickly as $\tilde{\omega}$ is further increased. We remind that in our CuCl$_{2}$ reference case $\tilde{E}$=0.195 a.u.=5.3 eV. \begin{figure} \vspace*{6cm} \special{psfile=FIG5.ps} \caption{The ratio $r(\protect\omega)$ between the satellite and the main peak for the divalent copper compounds CuBr$_2$, CuCl$_2$, and CuF$_2$. The parameters are given in Table II. The dotted lines are the limit values ($ r(\infty)$) for the respective cases. } \label{fig:Cu-dihalides} \end{figure} \subsection{Resolvent formulation} Alternatively, we can work in the energy space, and obtain the spectrum by direct inversion of a resolvent operator. We consider the Hamiltonian ${\cal % H}$, \begin{equation} {\cal H}={\cal H}_{0}+T+V, \label{eq:pt1} \end{equation} where by ${\cal H}_{0}$ we understand ${\cal H}_{0}(n_{c}=0)$. The exact final photoemission state $|\Psi _{f}^{sk}\rangle $ is\cite{Hedinbook} \begin{equation} |\Psi _{f}^{sk}\rangle =\left[ 1+\frac{1}{E-{\cal H}_{0}-T-V-i\eta }V\right] |\psi _{s}\rangle |\psi _{k}\rangle , \label{eq:pt2} \end{equation} where $|\psi _{s}\rangle $ ($s=1,2$) (Eq. (\ref{eq:s4})) are the exact (target) eigenstates of ${\cal H}_{0}(n_{c}=0)$ and $E=\epsilon _{k}+E_{s}$ is the energy of the final state. Using Eq. (\ref{eq:pt2}) we calculate the matrix element $M(s,{k})\equiv \langle \Psi _{f}^{sk}|\Delta |\Psi _{0}\rangle $ \begin{eqnarray} &&M(s,{k}) \label{eq:pt3} \\ &=&\langle \psi _{k}|\langle \psi _{s}|\left[ 1+V\frac{1}{\epsilon _{k}+E_{s}-{\cal H}_{0}-T-V+i\eta }\right] \Delta |\Psi _{0}\rangle . \nonumber \end{eqnarray} Introducing a basis set \begin{equation} |i\rangle =|\psi _{s}\rangle |\psi _{k}\rangle , \label{eq:d1} \end{equation} the matrix elements of $V$ can then be written as \begin{equation} V_{ij}\equiv V_{ks,k^{^{\prime }}s{^{\prime }}}=V_{kk^{^{\prime }}}v_{s}v_{s^{^{\prime }}}. \label{eq:d1a} \end{equation} Here \begin{equation} v_{s}=\cases{-{\rm sin}\varphi,& if $s=1$;\cr {\rm cos}\varphi,& if $s=2$,\cr} \label{eq:d1b} \end{equation} where we have used Eqs. (\ref{eq:8}, \ref{eq:s4}). The Hamiltonian matrix in this basis set is diagonalized, which gives the eigenvalues $\epsilon _{\nu }$ and the eigenvectors \begin{equation} |\nu \rangle =\sum_{i}c_{i}^{\nu }|i\rangle . \label{eq:d2} \end{equation} We then have \begin{eqnarray} &&M(s,k)\equiv M(i)= \label{eq:d3} \\ &&\langle i|\Delta |\Psi _{0}\rangle +\sum_{\nu }\sum_{j,l}{\frac{% V_{i,j}c_{j}^{\nu }c_{l}^{\nu }\langle l|\Delta |\Psi _{0}\rangle }{\epsilon _{k}+E_{s}-\epsilon _{\nu }+i\eta }}. \nonumber \end{eqnarray} The quantities $\langle i|\Delta |\Psi _{0}\rangle $ were given in Eq. (\ref {eq:s9}), $\langle i|\Delta |\Psi _{0}\rangle =m_{i}=m_{sk}=M_{k}w_{s}$. By organizing the sums in Eq. (\ref{eq:d3}) appropriately, the calculation of this expression is very fast and the main time is spent in diagonalizing the Hamiltonian matrix. We have found this method to be more efficient than the time-dependent method above. In the expression (\ref{eq:d3}), we can identify the first term as the intrinsic contribution, since this is the amplitude which is obtained if there is no interaction between the photoelectron and the target. The extrinsic effects are then determined by the square of the absolute value of the second term. The interference between the intrinsic and extrinsic contributions is given by the cross product of these terms. \subsection{Separable potential} It is interesting to consider a separable potential \begin{equation} V_{kk^{^{\prime }}}=\tilde{V}b_{k}b_{k^{^{\prime }}}, \label{eq:sep1} \end{equation} since it is then possible to obtain an analytical expression for $r(\omega )$. The operator in the denominator of Eq. (\ref{eq:pt2}) is written as \begin{equation} (z-{\cal H}_{0}-T-V)_{ij}=d_{i}(z)\delta _{ij}-\tilde{V}c_{i}c_{j}, \label{eq:sep2} \end{equation} where again $|i\rangle =|s\rangle |k\rangle $ is a combined index for the target state $s$ and the continuum state $k$ and $z$ is a (complex) number. Then \begin{equation} d_{sk}(z)\equiv d_{i}(z)=z-E_{s}-\epsilon _{k} \label{eq:sep3} \end{equation} and \begin{equation} c_{i}\equiv c_{sk}=b_{k}v_{s}. \label{eq:sep4} \end{equation} Using the fact that $V$ is separable, it is then straightforward to invert the expression in Eq. (\ref{eq:sep2}) and obtain \begin{eqnarray} &&[(z-{\cal H}_{0}-T-V)^{-1}]_{ij} \label{eq:sep5} \\ \nonumber &=&{\frac{\delta _{ij}}{d_{i}(z)}}+\tilde{V}{\frac{c_{i}c_{j}}{% d_{i}(z)d_{j}(z)(1-\tilde{V}\sum_{l}c_{l}^{2}/d_{l}(z))}}. \end{eqnarray} This leads to \begin{eqnarray} &&{\frac{r(\omega )}{r_{0}(\omega )}}= \label{eq:sep6} \\ &&\left| {\frac{1+{\rm cos}\varphi D_{k_{2}}(E_{0}+\omega )/{\rm cos}% (\varphi +\theta )}{1+{\rm sin}\varphi D_{k_{1}}(E_{0}+\omega )/{\rm sin}% (\varphi +\theta )}}\right| ^{2}, \nonumber \end{eqnarray} where $k_{s}$ is defined in Eq.~(\ref{eq:s10}), \begin{equation} D_{k}(\epsilon )=-{\frac{\tilde{V}}{\tilde{E}_{s}}}{\frac{b_{k}E(\epsilon )}{% M_{k}[1+(\tilde{V}/\tilde{E}_{s})C(\epsilon )]}}, \label{eq:sep7} \end{equation} with $\tilde{E}_{s}=1/(2\tilde{R}_{s}^{2})$ and \begin{equation} C(\epsilon )=-\tilde{E}_{s}\sum_{l}{\frac{c_{l}^{2}}{d_{l}(\epsilon +i\eta )}% } \label{eq:sep8} \end{equation} and \begin{equation} E(\epsilon )=-\tilde{E}_{s}\sum_{l}{\frac{c_{l}m_{l}}{d_{l}(\epsilon +i\eta )% }}. \label{eq:sep9} \end{equation} To obtain a model for $V_{kk^{^{\prime }}}$ we can, for instance, put \begin{equation} b_{k}=\sqrt{\frac{\tilde{R}_{s}}{R}}{\frac{(\tilde{R}_{s}k)^{2}}{1+(\tilde{R}% _{s}k)^{3}}}. \label{eq:sep10} \end{equation} Compared with the expression in Eq. (\ref{eq:ma6}), there is no term $% k-k^{^{\prime }}$ in the corresponding expression for $V_{kk^{^{\prime }}}$ . The neglect of this term means that $V_{kk^{^{\prime }}}$ goes to zero more slowly as one of the arguments $k$ or $k^{^{\prime }}$ goes to infinity. To compensate for this we use the power three for $\tilde{R}_{s}k$ in the denominator of (\ref{eq:sep10}), while in Eq. (\ref{eq:ma6}) the corresponding power is two. This is a reasonable approximation for small $k$% , but it breaks down for large $k$. \subsection{On the variables in the intensity ratio} For the satellite to main line intensity ratio we have, \[ r(\omega)=\frac{k_{1}}{k_{2}} \left|\frac{M(2,k_{2})}{M(1,k_{1})}% \right|^{2}. \] This ratio does not depend on any constant factor in $M_{k}$, since $M(s,k)$ is proportional to $M_{k}$. If we take the parameters $\tilde{R}_{d}$, $% \tilde{R}_{s}$, and $\tilde{R}_{sd}$ equal to a common typical radius $% \tilde{R}$ (as will be motivated later), and use the analytic expressions in Eqs. (\ref{eq:ma4}) and (\ref{eq:ma6}) then $r(\omega)$ or $% r(\omega)/r_{0}(\omega)$, apart from $\varphi$ and $\theta$, becomes a function of $\delta E/\tilde{E}$, $\tilde{V}/\tilde{E}$, and $\tilde{\omega }% /\tilde{E}$, with $\tilde{E}= (2\tilde{R}^{2})^{-1}$. We can see this since $% M_{k}$ is a function of $k\tilde{R}$, and $V_{kk^{\prime }}$ a function of $k% \tilde{R}$ and $k^{\prime }\tilde{R}$ apart from their prefactors. The prefactor of $V_{kk^{\prime}}$ is $\tilde{V}\tilde{R}/R$, while that for $% M_{k}$ has no influence. For each $V$ in a perturbation expansion of Eq. (% \ref{eq:pt3}) we have an energy denominator and a $k$-summation. The $k$% -summation gives an integral and a factor $Rdk.$ Using variables $\tilde{R}k$% , the $\tilde{R}/R$ in the prefactor vanishes. Factoring out $\tilde{E}$ in the energy denominator, we have a factor $\tilde{V}/\tilde{E}$ for each $% V_{kk^{\prime}}$, and instead of $\delta E$ and $\tilde{\omega}$ we have $% \delta E/\tilde{E}$ and $\tilde{\omega}/\tilde{E}$. With $\theta $ and $\varphi $ given, $\delta E$ is proportional to $U$. $U$ in turn is equal to $-V_{sc}(0) $, and thus somehow related to the strength of the scattering potential $\tilde{V}.$ If we fix the value of the sudden limit $r_{00}=\cot^{2}\left(\varphi+\theta\right)$ by choosing one of the angles, we still have an independent parameter left. This parameter can be used to decouple the relation between $\delta E$ and $\tilde{V}$ (whatever it is). Summarizing, we have found that the parameters of our model system appear as the angles $\theta $ and $\varphi$, and the excitation energy $% \delta E$ (or $U$), while the coupling between the photoelectron and the model system only appears in one parameter, $\tilde{V}/\tilde{E}=2\tilde{V}% \tilde{R}^{2}$, provided we use $\tilde{\omega }/\tilde{E}$ as variable. We have further motivated that we can vary the parameters $\delta E$ and $% \tilde{V}$ independently. \section{Approximate treatments} \subsection{Perturbation approach to lowest order in $V_{\lowercase{kk^{\prime}}}$} \label{sec:pert} The same problem can be also studied using the standard perturbation approach. We consider the expression for the matrix elements $M(s,k)$ in Eq. (\ref{eq:pt3}). To lowest order in $V$, we can neglect $V$ in the denominator of Eq. (\ref{eq:pt3}). Inserting the completeness relation $ \sum_{i}|i\rangle \langle i|=1$ in terms of eigenstates $|i\rangle \equiv |k\rangle |s\rangle $ we obtain \begin{eqnarray} &&M(s,k)=\langle s|\langle k|\Delta |\Psi _{0}\rangle \label{eq:pt0} \\ &&+\sum_{k^{^{\prime }}s^{^{\prime }}}V_{ks,k^{^{\prime }}s^{^{\prime }}}[(E-% {\cal H}_{0}-T+i\eta )^{-1}]_{k^{^{\prime }}s^{^{\prime }},k^{^{\prime }}s^{^{\prime }}}\langle s^{^{\prime }}|\langle k^{^{\prime }}|\Delta |\Psi _{0}\rangle . \nonumber \end{eqnarray} Using Eqs. (\ref{eq:d1a}, \ref{eq:d1b}) we obtain \begin{eqnarray} &M&(1,k)=-\sin (\varphi +\theta )M_{k} \nonumber \\ &-&\sin ^{2}\varphi \sin (\varphi +\theta )\sum_{k^{\prime }}\left[ \frac{% V_{kk^{\prime }}M_{k^{\prime }}}{E-E_{1}-\epsilon _{k^{\prime }}+i\eta }% \right] \nonumber \\ &-&\frac{\sin 2\varphi \cos (\varphi +\theta )}{2}\sum_{k^{\prime }}\left[ \frac{V_{kk^{\prime }}M_{k^{\prime }}}{E-E_{2}-\epsilon _{k^{\prime }}+i\eta }\right] , \label{eq:pt4} \end{eqnarray} \begin{eqnarray} &M&(2,k)=\cos (\varphi +\theta )M_{k} \nonumber \\ &+&\cos ^{2}\varphi \cos (\varphi +\theta )\sum_{k^{\prime }}\left[ \frac{% V_{kk^{\prime }}M_{k^{\prime }}}{E-E_{2}-\epsilon _{k^{\prime }}+i\eta }% \right] \nonumber \\ &+&\frac{\sin 2\varphi \sin (\varphi +\theta )}{2}\sum_{k^{\prime }}\left[ \frac{V_{kk^{\prime }}M_{k^{\prime }}}{E-E_{1}-\epsilon _{k^{\prime }}+i\eta }\right] , \label{eq:pt5} \end{eqnarray} where $V_{kk^{\prime }}=\langle k|V_{sc}|k^{\prime }\rangle $ and $% M_{k}=\langle k|\Delta |\psi _{c}\rangle $. We can then immediately calculate the photoemission spectra using Eq.~(\ref{eq:s8}). \begin{figure} \vspace*{8.5cm} \special{psfile=FIG6.ps} \caption{$r(\protect\omega)/r_0(\protect\omega)$ from the semi-classical approximation (SC), the first order perturbation expansion (PT) as well as the exact time evolution calculations for different values of the excitation energy $\protect\delta E$ and for $U/t=5.76$. The remaining parameters are taken from CuCl$_2$ ($R_0^{{\rm Cl}}=4.71$ a.u.). } \label{fig:ex_energy} \end{figure} \begin{figure} \vspace*{8.5cm} \special{psfile=FIG7.ps} \caption{The same as in Fig. \ref{fig:ex_energy} but varying the range $DR_0$ instead of $\protect\delta E$, where $D$ is a scale factor. The parameters of CuCl$_2$ are used. } \label{fig:range} \end{figure} In Figs. \ref{fig:ex_energy} and \ref{fig:range} we compare the perturbation expansion with the exact time dependent calculation for a realistic scattering potential in the symmetric case. In the symmetric case we have $% \delta E=2t\sqrt{1+r_{00}}= (U/2)\sqrt{1+1/r_{00}}$, and $U=-V_{sc}(0)$. Since the ratio $r_0(\omega)$ was discussed extensively in Sec. \ref {sec:sudden}, we here focus on $r(\omega)/r_0(\omega)$, which describes the effect of the scattering potential. We vary the excitation energy $\delta E$ by varying $t$, while keeping $r_{00}$ constant. We also vary the potential range by replacing $R_0$ in Eq.~(\ref{eq:m4}) by $DR_0$ and then varying $D$% . With $r_{00}$ fixed, $\delta E$ is proportional to $V_{sc}(0)$. Thus a small $\delta E$ and a small $D$ make the perturbation weak. The calculations are made for a range of parameter values around those given for CuCl$_2$ in Table \ref{used-parameter}. \subsection{Semi-classical approach} \label{sec:semi} We can also perform the photoemission calculation by assuming a classical trajectory of the emitted photoelectron\cite{Mahan}, producing a time-dependent potential which drives the dynamics of the the model. It has been reported that the semi-classical approach can give the unexpectedly good results for the systems with coupling to plasmons.\cite{Lars98,Inglesfield,Bardy} The essence of the semi-classical approach is to replace the scattering potential $V_{sc}(r)$ by a time-dependent potential using the charge density $\rho ({\bf r},\tau )$ of the emitted electron, {\it i.e.} \begin{equation} V_{sc}(r)\rightarrow \int d{\bf r}V_{sc}(r)\rho ({\bf r},\tau )=V_{sc}(v\tau ), \label{eq:c1} \end{equation} where we have used $\rho ({\bf r},\tau )=\delta ({\bf r}-{\bf v}\tau )$. We can then write the Hamiltonian as \begin{equation} {\cal H}(\tau )={\cal H}_{0}(n_{c}=0)+V(\tau ), \label{eq:c2} \end{equation} where ${\cal H}_{0}(n_{c}=0)$ can be expressed in terms of the exact final states ($\psi _{1}$ and $\psi _{2}$) in the presence of a core hole \begin{equation} {\cal H}_{0}(n_{c}=0)=E_{1}\psi _{1}^{\dagger }\psi _{1}+E_{2}\psi _{2}^{\dagger }\psi _{2}. \label{eq:c3} \end{equation} The time-dependent potential takes the form \begin{eqnarray}\label{eq:c4} &&V(\tau )=n_{b}V_{sc}(v\tau ) \nonumber \\ &=&V_{11}(\tau )\psi _{1}^{\dagger }\psi _{1}+V_{22}(\tau )\psi _{2}^{\dagger }\psi _{2}+V_{12}(\tau )(\psi _{1}^{\dagger }\psi _{2}+\psi _{2}^{\dagger }\psi _{1}), \nonumber \\ && \end{eqnarray} where (cf.~Eq. (\ref{eq:d1b})) \begin{eqnarray} \label{eq:c5} V_{11}(\tau ) &=&\sin ^{2}\varphi V_{sc}(v\tau ),\hskip0.3cmV_{22}(\tau )=\cos ^{2}\varphi V_{sc}(v\tau ), \nonumber \\ V_{12}(\tau ) &=&V_{21}(\tau )=-{\frac{1}{2}}\sin 2\varphi V_{sc}(v\tau ). \end{eqnarray} The remaining system (target) is still purely quantum mechanical, and we write its time-dependent wave function $|\Psi (\tau )\rangle $ as \begin{equation} |\Psi (\tau )\rangle =a_{1v}(\tau )|\psi _{1}\rangle e^{-iE_{1}\tau }+a_{2v}(\tau )|\psi _{2}\rangle e^{-iE_{2}\tau }. \label{eq:c6} \end{equation} The classical electron velocity $v$ is determined by energy conservation, that is, $\frac{1}{2}v^{2}=\tilde{\omega}$. We have here chosen the velocity corresponding to the satellite. We could also have performed two calculations, with the velocities corresponding to the leading peak and to the satellite, respectively. In contrast to the approach used here, this would, however, lead to the problem that the spectral weight would not be normalized. Applying the time-dependent Schr\"{o}dinger equation to $|\Psi (\tau )\rangle $, we obtain $a_{1v}(\tau )$ and $a_{2v}(\tau )$, \begin{equation} i\frac{\partial }{\partial \tau }a_{1v}(\tau )=V_{11}(\tau )a_{1v}(\tau )+V_{12}(\tau )a_{2v}(\tau )e^{-i\delta E\tau }, \label{eq:c7} \end{equation} \begin{equation} i\frac{\partial }{\partial \tau }a_{2v}(\tau )=V_{22}(\tau )a_{2v}(\tau )+V_{21}(\tau )a_{1v}(\tau )e^{i\delta E\tau }, \label{eq:c8} \end{equation} where $\delta E$ (Eq.~(\ref{eq:s7})) is the optical excitation energy. Eqs.~(% \ref{eq:c7}) and (\ref{eq:c8}) are subject to the initial conditions (cf Eq. \ref{eq:s9}) \begin{equation} a_{sv}(0)=w_{s}. \end{equation} The final photoemission currents $J_{i}(\omega )$ is \begin{equation} J_{i}(\omega )\propto |a_{iv}(\tau _{0})|^{2},\mbox{ }\mbox{ }i=1,2. \end{equation} It is sufficient to perform the calculation up to $\tau =\tau _{0}\equiv R_{0}/v$, since the potential vanishes for larger values of $\tau $. The relative intensity between main and satellite contributions is given by $% r(\omega )=J_{2}(\omega )/J_{1}(\omega )$ as before. In Figs. \ref{fig:ex_energy} and \ref{fig:range} we compare the semi-classical and exact results for a realistic potential in the symmetric case. The semi-classical theory is inaccurate over most of the energy range considered here. For large energies however, the semi-classical theory comes much closer to the exact result than does the PT. It is also clear that an increasing $\delta E$ ($\simeq V_{sc}(0)$) does not noticeably affect the energy for the adiabatic-sudden transition, where it strongly effects the maximum deviation. An increasing $D$ on the other hand not only strongly increases the maximum deviation, but also makes the adiabatic-sudden transition energy smaller. The dependence on the parameters will be investigated more extensively in the next section. \section{Adiabatic-sudden transition} \label{sec:adiabatic} We are now in a position to address the adiabatic-sudden transition and its dependence on the parameters. The calculations are performed with the analytical matrix elements in Eqs. (\ref{eq:ma4}, \ref{eq:ma6}). First we study the different characteristic lengths, $\tilde{R}_{d}$ for the dipole matrix elements, and $\tilde{R}_{s}$ and $\tilde{R}_{sd}$ for the scattering potential matrix elements. We find that it makes sense to use only one effective length $\tilde{R}$, and the corresponding energy $\tilde{E}=1/(2\tilde{R}^{2})$. As we remarked in Sec. VD, $r/r_{0}$ as a function of $% \tilde{\omega}/\tilde{E}$ depends on the parameters $\delta E/\tilde{E}$ and $\tilde{V}/\tilde{E}$, and also on the ''system'' parameters $\theta $ and $% \varphi $. We vary $\delta E$ independently of $\tilde{V}$, although for a given model there is a direct relation between these two quantities. Part of this relation can be offset by using different $\theta $ and $\varphi $ (with $r_{00}$ constant) but we do not explore this possibility. The exact solution with a separable potential is used to discuss the validity and breakdown of perturbation theory. We find that $\tilde{V}/\tilde{E}$ has a large effect on the deviation from the sudden limit, but little effect on the value of $\tilde{\omega}/\tilde{E}$ where the deviation becomes small, while $\delta E/\tilde{E}$ has a comparatively small effect on both magnitude and range of the deviation. For simplicity we use the CuCl$_{2}$ parameters $\theta=\varphi= 0.3$, which gives $r_{00}=2.1$. For CuCl$_2$ we further have $\tilde{V}$=-0.36 a.u., $\tilde{E}$=0.195 a.u. ($\tilde{R}$=1.6 a.u.), and $\delta E=0.237$ a.u., i.e., $\tilde V/\tilde E=-1.85$ and $\delta E/\tilde V=-0.66$. In our calculations, we vary $\tilde V/\tilde E$ and $\delta E/\tilde V$ by typically a factor of two around these reference values. \subsection{Exact numerical treatment with analytic matrix elements} We first illustrate the dependence on the ratio between the length scales $ \tilde{R}_{d}$, $\tilde{R}_{s},$ and $\tilde{R}_{sd}$. In Fig. \ref{fig:rd} we show the results for different ratios $\tilde{R}_{d}/\tilde{R}_{s}$ keeping $\tilde{R}_{sd}/\tilde{R}_{d}=1$. These results are obtained for $% \tilde{V}/\tilde{E}_{s}=-2.0$ and $\delta E/\tilde{V}=-0.5$, where $\tilde{E}% _{s}=1/(2\tilde{R}_{s}^{2})$ is the energy scale set by the scattering potential length scale. Fig. \ref{fig:rd} shows that as $\tilde{R}_{d}/% \tilde{R}_{s}$ is reduced the magnitude of the "overshoot" is increased. There are, however, no qualitative changes. \begin{figure} \centerline{\rotatebox{-90}{\epsfysize=3.5in \epsffile{FIG8.ps}}} \caption{The ratio $r(\protect\omega)/r_0(\protect\omega)$ as a function of $ \tilde R_d/\tilde R_s$ for a fixed $\tilde R_{sd}/\tilde R_s=1$ and for $\protect\varphi=\protect\theta=0.3$. The figure illustrates that there are no qualitative changes as the length scales for the dipole and scattering matrix elements become different.} \label{fig:rd} \end{figure} Fig. \ref{fig:rsd} shows results for different values of $\tilde{R}_{sd}/ \tilde{R}_{s}$ for a fixed $\tilde{R}_{d}/ \tilde{R}_{s}=1$. From Eq. (\ref {eq:ma6}) it can been seen that this corresponds to varying the range of values $k-k^{^{\prime }}$ where $V_{kk^{^{\prime }}}$ is large, without changing the range over which $V_{kk}$ varies. The figure illustrates that the overshoot becomes larger as $\tilde{R}_{sd}/\tilde{R}_{s}$ is reduced. This is natural, since decreasing $\tilde{R}_{sd}$ effectively makes the scattering potential stronger by expanding the range of values $ k-k^{^{\prime }}$ with large scattering matrix elements. The qualitative behaviour, however, is not changed. In view of Figs. \ref{fig:rd} and \ref {fig:rsd}, we study below the case when $\tilde{R}_{sd}= \tilde{R}_{s}=\tilde{ R}_{d}$, as mentioned in Sec. VD. \begin{figure} \centerline{\rotatebox{-90}{\epsfysize=3.5in \epsffile{FIG9.ps}}} \caption{The ratio $r(\protect\omega)/r_0(\protect\omega)$ as a function of $\tilde R_{sd}/\tilde R_s$ for a fixed $\tilde R_{d}/\tilde R_s=1$ and for $\protect\varphi=\protect\theta=0.3$. The figure illustrates that there is no qualitative changes as the ratio of the two length scales in the scattering matrix elements is varied.} \label{fig:rsd} \end{figure} Fig. \ref{fig:strength} shows such results for different values of the strength of the scattering potential $\tilde V/\tilde E$ and for different values of the excitation energy $\delta E/\tilde V$. In each panel $\delta E /\tilde V$ is kept fixed, but the ratio is varied by a factor of four from Fig. \ref{fig:strength}a to Fig. \ref{fig:strength}c. Typically $r(\omega)/r_0(\omega)$ has an overshoot for small values of $\tilde{\omega}$. For somewhat larger $\omega$ the ratio approaches unity and possibly becomes smaller than unity. The overshoot can be fairly large and happens on a small energy scale ($\sim \tilde E$). In a few cases of a large overshoot, $r(\omega)/r_0(\omega)$ does not become approximately unity until $\tilde{\omega}$ is several times $\tilde E$, although the relevant energy scale is still $\tilde E$. In the case of an undershoot, $r(\omega)/r_0(\omega)$ approaches unity from below very slowly (energy scale much larger than $\tilde E$). The undershoot is, however, relatively small, and if we do not require a high accuracy, we consider the sudden approximation is valid when the overshoot becomes small. This means that as the range of the scattering potential is made larger, the sudden limit is reached at a smaller energy. This is the opposite to what one would expect from the semi-classical theory. The figure illustrates that $\delta E$ is not the relevant energy scale. Since in each panel we keep $\delta E/% \tilde V$ fixed, there is a variation of $\delta E/\tilde{E}$ by a factor of four. Furthermore there is a variation of $\delta E/\tilde{V}$ by a factor of four in going from the top to the bottom panel in Fig. \ref{fig:strength}. There is no corresponding change in the energy for the adiabatic to sudden transition. \subsection{Separable potential} It is interesting to study a separable potential, since it is then possible to obtain an analytical solution. This makes it easier to interpret the results. It also allows the study the effects of multiple scattering, i.e. the deviations from first order perturbation theory. Fig. \ref{fig:separable} shows results of the exact and first order theory using the same values of $ \delta E/\tilde{V}$ and $\tilde{V}/\tilde{E}$ as in Fig. \ref{fig:strength}b. The separable potential overestimates the magnitude of the overshoot in $r(\omega )/ r_{0}(\omega)$ quite substantially. Otherwise the results are rather similar. For a qualitative discussion, we can therefore use the separable potential. \begin{figure} \centerline{\rotatebox{-90}{\epsfxsize=2.5in \epsffile{FIG10.ps}}} \vskip0.8cm \caption{The ratio $r(\protect\omega )/r_{0}(\protect\omega )$ as a function of $\tilde{\protect\omega}/\tilde{E}$ for different values of $\tilde{V}/% \tilde{E}$ and $\protect\delta E/\tilde{V}$ and for $\protect\varphi =% \protect\theta =0.3$. The figure illustrates that $\tilde{E}$ is an appropriate energy scale for the adiabatic to sudden transition.} \label{fig:strength} \end{figure} For simplicity, we consider $\delta E=0$. We further put $M_{k}=(\tilde{R}k)^{2}/\left[ 1+(\tilde{R}k)^{3}\right] =b_{k}$. This is a poor approximation for large $k$, but then anyhow also $V_{kk^{\prime }}$ is poorly represented by the separable potential. Our approximations lead to simple results for the functions $C$, $D$ and $E$ entering in Eqs. (\ref{eq:sep6}-\ref{eq:sep10}). \begin{equation} C(\epsilon )=-\tilde{E}\sum_{k^{^{\prime }}}{\frac{b_{k^{^{\prime }}}^{2}}{% \epsilon -\epsilon _{k^{^{\prime }}}+i\eta }} \label{eq:as1} \end{equation} and $E(\epsilon )={\rm cos}\theta \tilde{M}C(\epsilon )$. Then \begin{equation} D(\epsilon )=-{\rm cos}\theta {\frac{\tilde{V}}{\tilde{E}}}{\frac{C(\epsilon )}{1+(\tilde{V}/\tilde{E})C(\epsilon )}}. \label{eq:as2} \end{equation} Since $D_{k}$ is independent of $k$ in this approximation, we have dropped the index $k$. The function $C(\epsilon )$ is shown in Fig. \ref {fig:universal}. For $\varphi =\theta $ we can then write \begin{equation} {\frac{r(\omega )}{r_{0}(\omega )}}=\left| {\frac{1-\frac{{\rm cos}% ^{2}\varphi }{{\rm cos}2\varphi }\frac{\tilde{V}}{\tilde{E}}C(\tilde{\omega}% )/[1+\frac{\tilde{V}}{\tilde{E}}C(\tilde{\omega})]}{1-\frac{1}{2}\frac{% \tilde{V}}{\tilde{E}}C(\tilde{\omega})/[1+\frac{\tilde{V}}{\tilde{E}}C(% \tilde{\omega})]}}\right| ^{2}. \label{eq:as3} \end{equation} For a level ordering as indicated in Fig. \ref{fig:schematic}, $0<\varphi <\pi /4$ and cos$^{2}\varphi /{\rm cos}(2\varphi )\geq 1$. In our standard case with $\varphi =0.3$, $\cos ^{2}\varphi /\cos (2\varphi )=1.11$. Thus the term in the numerator of Eq. (\ref{eq:as3}) dominates. The factor $1+(% \tilde{V}/\tilde{E})C(\tilde{\omega})$ gives the multiple scattering, which is not included in the first order perturbation theory, i.e. the first order result is \begin{equation} \left[ \frac{r(\omega )}{r_{0}(\omega )}\right] _{{\rm PT}}=\left| \frac{1-% \frac{\cos ^{2}\varphi }{\cos 2\varphi }\frac{\tilde{V}}{\tilde{E}}C(\tilde{% \omega})}{1-\frac{1}{2}\frac{\tilde{V}}{\tilde{E}}C(\tilde{\omega})}\right| ^{2}. \label{eq:as3a} \end{equation} We now compare the behaviors of Eqs. (\ref{eq:as3}) and (\ref{eq:as3a}), to see the effects of using perturbation theory. For small values of $\tilde{\omega}$, Re$C(\tilde{\omega})$ is positive and then changes sign at about $\tilde{ \omega}/\tilde{E}\sim 2$. Im $C(\tilde{\omega})$ is always positive. Due to our crude approximations for $b_{k}$ and $M_{k}$, $C(\tilde{\omega})$ rapidly becomes unreliable beyond $\tilde{\omega}/\tilde{E}=2$. Both for the exact and perturbative expressions $r/r_{0}$ goes from over- to undershoot approximately when Re$C(\tilde{\omega})=0$. This is somewhat earlier than in Fig. 11, where however $\delta E=0.5$. Comparing $r(\omega)/r_{0}(\omega )$ for the exact (Eq. (\ref{eq:as3})) and the first order result (Eq. (\ref {eq:as3a})), we find that the exact solution is larger when ${\rm Re}C(% \tilde{\omega})\gtrsim {\rm Im}C(\tilde{\omega})$, cf Fig. \ref {fig:separable}. This is consistent with second order perturbation theory, which is found to enhance $r(\omega )/r_{0}(\omega )$ for small $\tilde{ \omega}$ and reduce it for large $\tilde{\omega}$. For $\tilde{\omega}=0$, $C(\tilde{\omega})$ is purely real and slightly larger than 0.1, thus multiple scattering gives a divergence in both $J_{1}(\tilde{\omega})$ and $J_{2}(\tilde{\omega})$ when $\tilde{V}/\tilde{E}\sim -10$. This is due to the attractive potential $V$ forming a bound state from the continuum states. \begin{figure} \vspace*{9.5cm} \special{psfile=FIG11.ps} \caption{The ratio $r(\protect\omega)/r_0(\protect\omega)$ for a separable scattering potential (\ref{eq:sep1}). According to the exact result in upper panel, the overshoot is substantially overestimated by the separable potential compared to Fig. \ref{fig:strength}b, but there is still a qualitative agreement with the more realistic model (\ref{eq:ma6}). The lower panel shows how perturbation theory works and it illustrates the effects of multiple scattering. } \label{fig:separable} \end{figure} The important conclusion from analysing the separable potential is that if $\tilde V$ is not too large, first order perturbation theory gives roughly the correct range over which there are essential deviations from the sudden limit, while multiple scattering increases the magnitude of these deviations for small $\tilde{\omega}$ and slightly decreases them for larger $\tilde{\omega}$. \begin{figure} \vspace*{9.5cm} \special{psfile=FIG12.ps} \caption{The function $C(\epsilon)$ relevant for a separable potential is given in the upper panel. The low panel gives the behaviors of $F(\epsilon)$ defined in Eq.(\ref{eq:Fk1}). The dotted lines give the asymptotic behaviors of $F(\epsilon)$ in Eq.(\ref{eq:apt8}). } \label{fig:universal} \end{figure} We next discuss the physical interpretation of the expression (\ref{eq:as3}) (or the perturbational expressions in (\ref{eq:pt0}, \ref{eq:pt4}, \ref{eq:pt5})). Here unity (the first term in (\ref{eq:pt0}, \ref{eq:pt4}, \ref{eq:pt5})) corresponds to a direct transition into the final continuum state corresponding to energy conservation. The second term (the last two terms in (\ref{eq:pt0}, \ref{eq:pt4}, \ref{eq:pt5})) corresponds to a virtual transition into some other continuum state followed by one or several scattering events with the electron ending up in the continuum state corresponding to energy conservation. Let us consider the virtual emission into a continuum state with a larger energy than the final state and let this be followed by one scattering event into the final state. For a negative $\tilde V$ the interference with the direct event is then constructive. For small photon energies such events dominate for two reasons. Firstly, there are many more states available above the energy corresponding to energy conservation than below, and secondly the dipole matrix elements suppress the transitions to the energies below. As a result, both the main peak and the satellite are enhanced by the scattering effects. For the values of $\varphi$ and $\theta$ considered here ($<\pi/4$), the relative effect is stronger for the satellite. As a result $r(\omega)/r_0(\omega)$ is enhanced. For larger photon energies Re$C(\tilde{\omega})$ becomes negative. The density of states of partial waves with given $l$ and $m$ quantum numbers decreases with energy ($\sim 1/\sqrt{\epsilon})$). This favours virtual emissions to states below the final continuum state. Depending on the model for $b_k$ and $M_k$, these matrix elements may have the same effect. As a result, Re $C(\tilde{\omega})$ becomes slightly negative for large energies and the ratio $r(\omega)/r_0(\omega)$ is slightly smaller than one. The relevant energy scale for $C(\omega)$ is $\tilde E$. This is a combination of the two effects discussed above. The turn on of the dipole matrix elements on an energy scale of the order $\tilde E$ favors an increasing value of $C$ over this energy scale, while the density of states effects becomes more important for larger energies. As a result, both Re $C(\omega)$ and Im $C(\omega)$ have a maximum at an energy of the order $\tilde E$. \subsection{Perturbational treatment with analytic matrix elements} In this section we study the perturbation theory expression in more detail and without relying on a separable potential. Instead we consider the more realistic matrix elements in Eqs. (\ref{eq:ma4}) and (\ref{eq:ma6}), assuming that $\tilde{R}_{d}=\tilde{R}_{s}= \tilde{R}_{sd}=\tilde{R}$. We define a function $F_{k}$ by \begin{equation} \sum_{k^{^{\prime }}}{\frac{V_{kk^{^{\prime }}}M_{k^{^{\prime }}}}{\epsilon -\epsilon _{k^{^{\prime }}}+i\eta }}\equiv -{\frac{\tilde{V}}{\tilde{E}}}% M_{k}F_{k}(\epsilon /\tilde{E}), \label{eq:Fk1} \end{equation} which is possible due to the simple form of $V_{kk^{^{\prime }}}$ and $% M_{k^{^{\prime }}}$. Explicitly we have \begin{equation} F_{k}\left( \epsilon \right) =\frac{1}{\pi }\int_{0}^{\infty }\frac{% x^{4}dx}{\left[ 1+x^{2}\right] ^{2}\left[ 1+\left( \tilde{R}k-x\right) ^{2}\right] \left[ x^{2}-\epsilon -i\eta \right] }. \label{eq:Fk2} \end{equation} From Eqs. \ref{eq:pt4} and \ref{eq:pt5} we have \begin{eqnarray}\label{eq:as4} & &\frac{r(\omega )}{r_{0}(\omega )}= \\ \nonumber & &\left| \frac{% 1-\cos ^{2}\varphi {\frac{\tilde{V}}{\tilde{E}}}F_{k_{2}}\left(\frac{\tilde{\omega }}% {\tilde{E}}\right)-\frac{\sin 2\varphi \sin (\varphi +\theta )}{2\cos (\varphi +\theta )}{\frac{\tilde{V}}{\tilde{E}}}F_{k_{2}}\left(\frac{ \tilde{\omega } +\delta E}{\tilde{E}}\right)}{1-\sin ^{2}\varphi {\frac{\tilde{V}}{\tilde{E}}% }F_{k_{1}}\left(\frac{\tilde{\omega }+\delta E} {\tilde{E}}\right)-\frac{\sin 2\varphi \cos (\varphi +\theta )}{2\sin (\varphi +\theta )}{\frac{\tilde{V}}{ \tilde{E}}}F_{k_{1}}\left(\frac{\tilde{\omega }} {\tilde{E}}\right)}\right| ^{2}. \end{eqnarray} Since $\tilde{R}k_{2}=\sqrt{\tilde{\omega }/\tilde{E}}$ and $% \tilde{R}k_{1}=\sqrt{\left( \tilde{\omega }+\delta E\right) /\tilde{E% }}$ we see, as stated earlier, that $r/r_{0}$ depends only on $\tilde{V}/% \tilde{E}$, $\delta E/\tilde{E}$ and $\tilde{\omega }/\tilde{E}.$ First we consider the limit of small values of $k$ and $\delta E$. For $ \tilde{\omega }=\delta E=0$ we have $F_{0}\left( 0\right) =1/16$, and \begin{equation} \frac{r(0)}{r_{0}(0)}=\left[ {\frac{1+\frac{|\tilde{V}|}{\tilde{E}% }\cos \varphi \cos \theta /(16\cos (\varphi +\theta ))}{1+\frac{ |% \tilde{V}|}{\tilde{E}} \sin \varphi \cos \theta /(16\sin (\varphi +\theta ))}}\right] ^{2}. \label{eq:apt7} \end{equation} If the more localized level $a$ is above $b$ (see Fig. \ref{fig:schematic}) in the initial state and below $b$ in the final state (``shake-down''), we have $0<\varphi <\pi /4$ and $0<\theta <\pi /4$, and the factor in the brackets is larger than unity. Thus interaction effects enhance the ratio $r(\omega )/r_{0}(\omega )$. This corresponds to a constructive interference between intrinsic and extrinsic effects. This is in contrast to the destructive interference found for the plasmon case.\cite{Lars98,plasmon} The present treatment, however, refers to the ``shake-down'' scenario, and it is more appropriate to compare the plasmon case with the ''shake up'' case ($-\pi /4<\varphi <0<\theta <\pi /4$ and $\varphi +\theta <0$). Then the expression (\ref{eq:apt7}) for $r(\omega )/r_{0}(\omega )$ indeed becomes smaller than one, and the relative weight of the satellite is reduced for small energies. We notice, however, that both the satellite and the main peak are enhanced by the interference, but that the main peak is enhanced more in the ''shake-up'' situation. We next consider the case when $k$ is large. $F_{k}\left( \epsilon \right) $ for large $k$ and $\epsilon $ is (with $\tilde{R }k\approx \sqrt{\epsilon }$) \begin{equation} F_{k}\left( \epsilon \right) =\frac{1}{2\sqrt{\epsilon }}\left[ i- \frac{1}{2\sqrt{\epsilon }}+\tilde{R}k-\sqrt{\epsilon}\right]. \label{eq:apt8} \end{equation} For the case when $\theta =\varphi $ we obtain \begin{equation} \frac{r(\omega )}{r_{0}(\omega)}-1=-{\frac{|\tilde{V }|}{2\tilde{\omega }}\frac{1+\delta E/\tilde{E}}{2\cos (2\varphi )}+ \frac{|\tilde{V}|^{2}}{4\tilde{\omega }\tilde{E}}}\left( {\frac{\cos^{4}\varphi }{\cos^{2}(2\varphi )}}-{\frac{1}{4}}\right). \label{eq:apt8a} \end{equation} Thus the approach to the sudden limit goes as $1/\tilde{\omega }$ with a coefficient which depends on the parameters. With our standard CuCl$_{2}$ parameters, we have \[ \frac{r(\omega )}{r_{0}(\omega )}-1=-0.30{\frac{|% \tilde{V}|}{\tilde{\omega }}}\left( 1+\frac{\delta E}{\tilde{E}}% \right) +0.24{\frac{|\tilde{V}|^{2}}{\tilde{\omega }\tilde{E}}} \] We note that for large $|\tilde{V}|/\tilde{E}$, and when $|\tilde{V}|$ is large enough compared to $\delta E,$ the last (positive) term dominates. The approach to the sudden limit is then set by $|\tilde{V}|^{2}/\tilde{E }=2\left( \tilde{R} \tilde{V}\right) ^{2}$. To evaluate Eq.(\ref{eq:as4}) when $\delta E=0$ we only need the function $F(\epsilon)$, $$ F\left( \epsilon \right) =\frac{1}{\pi }\int_{0}^{\infty }\frac{% x^{4}dx}{\left[ 1+x^{2}\right] ^{2}\left[ 1+\left( \sqrt{\epsilon}-x\right) ^{2}\right] \left[ x^{2}-\epsilon -i\eta \right] }. $$ We show $F\left( \epsilon \right) $ in the lower panel of Fig.\ref{fig:universal}. The results Eq. (\ref{eq:apt8}) for large values of $\epsilon $ are shown by the dotted lines in the figure. Clearly the approach of Re $% F$ to its asymptote is very slow. If we take $\delta E=0$ we have the same form as in first order perturbation theory with a separable potential Eq. (\ref{eq:as3a}), \begin{equation} \frac{r(\omega )}{r_{0}(\omega )}=\left| \frac{1-\frac{\cos ^{2}\varphi }{% \cos 2\varphi }\frac{\tilde{V}}{\tilde{E}}F(\tilde{\omega})}{1-\frac{1}{2}% \frac{\tilde{V}}{\tilde{E}}F(\tilde{\omega})}\right| ^{2}, \label{eq:apt100} \end{equation} where we have put $\varphi=\theta$. As shown in Fig. \ref{fig:universal}, $F(\epsilon )$ has a qualitatively similar behavior as $C(\epsilon )$ for $\epsilon/\tilde E\lesssim 2$. As in the case of the function $C$, the relevant energy scale is $\tilde E$. \begin{figure} \vspace*{1.cm} \centerline{\rotatebox{-90}{\epsfxsize=2.5in \epsffile{FIG13.ps}}} \vskip0.8cm \caption[]{\label{fig:pert} The ratio $r(\omega)/r_0(\omega)$ as a function of $\tilde{\omega}/\tilde E$ for different values of $\tilde V/\tilde E$ and $\delta E/\tilde V$.} \end{figure} In Fig. \ref{fig:pert} we show $r(\omega )/r_{0}(\omega )$ as a function of $% \tilde{\omega}/\tilde{E}$ for a few values of $\delta E/\tilde{V}$ and $% \tilde{V}/\tilde{E}$. We see that $r(\omega )/r_{0}(\omega )$ starts at a positive value (cf Eq.~(\ref{eq:apt7})), and reaches a broad maximum at about $\tilde{\omega}/\tilde{E}\sim 0.5-1.5$. Compared to the separable potential solution in Fig. \ref{fig:separable}, the overshoot behavior is robust up to fairly large energies, which is due to Im$F(\tilde{\omega})$ decaying more slowly than Im$C(\tilde{\omega})$. For much larger values of $% \omega ,$ $\left( r/r_{0}-1\right) $ decays as $1/\tilde{\omega}$, as shown in Eq.~(\ref{eq:apt8a}). Here, based on the discussion in Sec. VIIB, we can expect that as multiple scattering becomes important, the region where there is an overshoot is substantially reduced and the region with an undershoot becomes larger. At the same time, the overshoot intensity will be enhanced. These behaviors are actually confirmed by comparing with the exact calculations given in Fig. \ref{fig:strength}. Fig. \ref{fig:pert} illustrates that for intermediate values of $\tilde{% \omega}$, when Re$F$ dominates, $\left( r/r_{0}-1\right) $ goes as roughly $% |\tilde{V}|/\tilde{\omega}$ if $\delta E/\tilde{E}$ is not too large. For larger (but not too large) values of $\tilde{\omega}$, Re$F$ becomes small and Im$F$ dominates. Since Re$F$ is positive for small energies, this leads to a constructive interference between intrinsic and extrinsic effects. For energies of the order $\tilde{E}$, Re$F$ changes sign, and the interference becomes weakly destructive. For somewhat larger energies the extrinsic effects are mainly determined by the imaginary part of $F$. From Eq. (\ref{eq:apt8a}) it follows that in perturbation theory the extrinsic effects become small on the energy scale $\tilde{V}^{2}/\tilde{E}$. \subsection{Semi-classical approximation} In this section we analyze the adiabatic-sudden transition within the semi-classical framework. From the coupled differential equations Eqs.~(\ref {eq:c7}) and (\ref{eq:c8}) we can obtain differential equations for $% \partial |a_{iv}(\tau )|^{2}/\partial \tau $, $i=1,2$. Integration of these equations, leads to \begin{eqnarray}\label{eq:a1} &&|a_{2v}(\tau _{0})|^{2}-|a_{2v}(0)|^{2} \nonumber \\ &=&2{\rm Im}\int_{0}^{\tau _{0}}V_{12}(\tau )a_{1v}(\tau )a_{2v}^{\ast }(\tau )e^{i\delta E\tau }d\tau , \end{eqnarray} where $\tau _{0}=R_{0}/v$ is the time at which the emitted electron with the velocity $v$ leaves the range $R_{0}$ of the scattering potential. $% |a_{iv}(\tau _{0})|^{2}-|a_{iv}(0)|^{2}$ is a measure of the deviation from the sudden limit. For small values of $\tau $ both the coefficients $% a_{iv}(\tau )$ and the exponent $e^{i\delta E\tau }$ are approximately real and there is a small contribution to the imaginary part of the integral in Eq.~(\ref{eq:a1}). As $\tau $ grows there is, however, a contribution from both these sources. To obtain a qualitative understanding of the semi-classical approximation, we solve the Schr\"{o}dinger equations Eqs.~(\ref{eq:c7}) and (\ref{eq:c8}) to lowest order in $1/v$. This leads to \begin{eqnarray} &&a_{1v}(\tau )=a_{1v}(0) \label{eq:a1a} \\ &&-i\int_{0}^{\tau }d\tau ^{^{\prime }}[V_{11}(\tau ^{^{\prime }})a_{1v}(0)+V_{12}(\tau ^{^{\prime }})a_{2v}(0)] \nonumber \end{eqnarray} and a similar result for $a_{2v}(\tau )$. This gives \begin{eqnarray} &&|a_{2v}(\tau _{0})|^{2}-|a_{2v}(0)|^{2}={\frac{1}{2}}\sin (2\varphi )\left\{ {\frac{1}{2}}\sin (2\theta )\right. \label{eq:a1b} \\ &&\left. \times \left[ \int_{0}^{\tau _{0}}d\tau V_{sc}(\tau )\right] ^{2}+\sin (2\varphi +2\theta )\delta E\int_{0}^{\tau _{0}}d\tau \tau V_{sc}(\tau )\right\} . \nonumber \end{eqnarray} To discuss the result, we for a moment assume a simple $\tau $-dependence of $V_{ij}(\tau )$ \begin{equation} V_{ij}(\tau )=V_{ij}(0)(1-{\frac{\tau }{\tau _{0}}}), \label{eq:a3} \end{equation} which corresponds to the $r$-dependence used in Eq.~(\ref{eq:ma6a}). We note, however, that this form is too simple to describe the behavior of the more realistic potential in Eq.~(\ref{eq:m4}). Inserting Eqs.~(\ref{eq:a1a}) and (\ref{eq:a3}) in Eq.~(\ref{eq:a1b}) gives \begin{eqnarray}\label{eq:a4} \Delta _{2} &\equiv |&a_{2v}(\tau _{0})|^{2}-|a_{2v}(0)|^{2}={\frac{1}{4}}% V_{sc}(0)\sin (2\varphi ) \nonumber \\ &&\times \left[ {\frac{1}{4}}\sin (2\theta )V_{sc}(0)+{\frac{1}{3}}\sin (2\varphi +2\theta )\delta E\right] \tau _{0}^{2}. \end{eqnarray} We now extend this treatment to intermediate values of $v$ where the adiabatic to sudden transition takes place. Using the expressions Eqs.~(\ref {eq:s2}), (\ref{eq:s5}) and (\ref{eq:s7}) to relate $\delta E$ and $% U=-V_{sc}(0)$ we obtain \begin{equation} \frac{r(\omega )}{r_{00}}-1=\frac{\Delta _{2}}{\sin ^{2}(2\varphi )\cos ^{2}(2\varphi )}=-{\frac{\sin (2\varphi )}{\sin (2\theta )}}{\frac{(\delta E)^{2}\tau _{0}^{2}}{12}}. \label{eq:a5} \end{equation} {\it Within the semi-classical theory} the condition for the sudden approximation is then \begin{equation} {\frac{v}{R_{0}}}={\frac{1}{\tau _{0}}}\gg \delta E\sqrt{\frac{\sin (2\varphi )}{12\sin (2\theta )}}. \label{eq:a6} \end{equation} We are now in a position to discuss the approach to the sudden limit. Within a semi-classical framework it seems clear that we have to require that the hole potential is fully switched on after a ``short'' time $\tau _{0}={% v/R_{0}}$, i.e., that the emitted electron leaves the range of the scattering potential after a short time. The question is, however, what we mean by ``short''. From Eq.~(\ref{eq:a4}) it follows that the time-scale is set by both the inverse of $\delta E$ and the inverse of $V_{sc}(0)$. In Eq.~(\ref{eq:a6}) we have used the relation between $\delta E$ and $V_{sc}(0) $ to remove $V_{sc}(0)$ from Eq.~(\ref{eq:a6}). From Eq.~(\ref{eq:a6}) we obtain the condition for the sudden approximation within the SC theory \begin{equation} \label{eq:a6a} \tilde{\omega}={\frac{1}{2}}k^2\gg (\delta E)^2 R_0^2={\frac{(\delta E)^2}{2% \tilde E}}\sim {\frac{V_{sc}(0)^2 }{\tilde E}}. \end{equation} Thus, according to the semi-classical theory, the sudden approximation requires that $\epsilon_k\gg(\delta E)^2/\tilde E$. Comparison with the full quantum mechanical calculations in Fig. \ref{fig:strength} shows that this criterion is not appropriate for the range of parameters considered here. The reason is that we have considered a parameter range where the semi-classical theory is not very accurate. It is interesting that the SC theory correctly predicts that the weight of the satellite goes to zero at threshold. Nevertheless, the SC theory does not give the correct physics at the threshold. In the full quantum mechanical calculation the weight of the satellite goes to zero due to the effects of the dipole matrix element, which becomes very small at small photoelectron energies. This effect is not included in the SC theory. In the semi-classical treatment, the small weight of the satellite is due to the fact that the scattering potential between the outgoing slow electron and the excitation means that the hole potential is only switched on slowly. In the quantum mechanical treatment, on the other hand, the scattering potential leads to an enhancement of the relative weight of the satellite close to the threshold for the shake-down case. \section{Discussion} We have studied the photoemission spectrum of a simple model with a localized charge transfer excitation. We have obtained exact numerical results for the spectrum as a function of the photon energy $\omega$ and in particular focussed on the ratio $r(\omega)$ between the weights of the satellite and the main peak. These calculations are compared with perturbational and semi-classical treatments. The results have been analyzed using the latter two approaches. An important effect in the ratio $r(\omega)$ is due to the energy dependence of the dipole matrix elements and a factor $1/(\partial \epsilon_k/\partial k)\sim 1/k$ in the expression for the spectrum. This leads to a suppression of the satellite close to the threshold, but can lead to an overshoot further away from the threshold. This effect was discussed in Sec. \ref {sec:sudden} and is described by $r_0(\omega)$. If the interaction between the emitted electron and the target is weak, this effect dominates. It is determined by the excitation energy $\delta E$ and the relevant energy scale $\tilde E_d$ of the dipole matrix element. If $\delta E/\tilde E_d$ is small, $ r_0$ reaches its limiting value from below, while there is an overshoot if $\delta E/\tilde E_d\gg 1$. In both cases $r_0$ reaches its limiting value for a kinetic energy of the order a few times $\delta E$. To study the effects of the scattering potential between the emitted electron and the target we have focussed on the ratio $r(\omega)/r_0(\omega)$. This quantity shows an overshoot for small values of $\omega$ in the ``shake-down'' situation studied here. Depending on the parameters there may be an undershoot for larger energies, which extends over a large energy range. This undershoot is, however, fairly small for the cases considered here. The sudden approximation is then valid to a reasonable accuracy when the overshoot has become small. We show that this happens on the energy scale $\tilde E=1/(2\tilde R^2)$, where $\tilde R$ is a typical length scale of the scattering potential. One of the main results of this paper is that for a coupling to localized excitations, the adiabatic to sudden transition takes place at quite small kinetic energies of the photoelectron. This is in contrast to the large kinetic energies needed for the case of coupling to plasmons. In the plasmon case, the kinetic energy is typically so large that the semi-classical treatment is a very good approximation. The adiabatic to sudden transition is then expected to happen on the energy scale $(\omega_q\lambda)^2$,\cite{Inglesfield} where $\omega_q$ and $\lambda$ are the plasmon frequency and wavelength, respectively. Since the long wavelength plasmons dominate the transition, this happens at very large energies. For a localized excitation, the relevant length scale of the scattering potential is smaller, and the transition is expected to take place at a smaller energy scale. Actually, the transition takes place at such a small energy that the semi-classical theory is usually not valid any more. It is interesting that the semi-classical theory therefore predicts the opposite dependence on the range $\tilde R$ of the scattering potential, namely as $\tilde R^2$ instead of $\tilde E =1/(2\tilde R^2)$. For the ''shake-down'' scenario considered here (the two outer levels cross as the hole is created), we find constructive interference (increase of $r(\omega )/r_{0}(\omega )$) between the intrinsic and extrinsic processes at low photoelectron energies. This is in contrast to the destructive interference found in the plasmon case and to the reduction of $r(\omega )/r_{0}(\omega )$ found here for the ''shake-up'' case (no level crossing). \section*{acknowledgement} This work has been supported by the Max-Planck Forschungspreis. One of us (LH) carried out part of his contribution to this work at the Max-Planck Institute for Festk\"{o}rperforschung.
\section{Introduction} The HEGRA system \cite{hegra_perf,hegra_trigger,hegra_mkn} of imaging atmospheric Cherenkov telescopes (IACTs) is the first installation employing the stereoscopic observation of air showers with multiple Cherenkov telescopes on a routine basis. Compared to individual telescopes, IACT stereoscopy provides an improved reconstruction of shower parameters and better background rejection (see, e.g., \cite{stereo}). Major new instruments for VHE gamma-ray astronomy now in the construction phase -- such as VERITAS \cite{veritas} and HESS \cite{hess} -- are based on the concept of IACT stereoscopy, with frequently half a dozen or more telescopes observing the same shower from different viewing angles. In Cherenkov telescopes \cite{iact_review}, the Cherenkov light emitted by shower particles is imaged onto a ``camera'' in the focal plane of a large reflector, generating an elongated, roughly elliptical image. The major axis of the image represents the image of the shower axis. Therefore, the major axis of the image points towards the image of the source on one side, and to the point where the shower axis intersects the plane of the telescope dish on the other side (Fig.~\ref{fig_image}). If a shower is observed by a stereoscopic system of two Cherenkov telescopes, its direction (i.e., the image of the source) can be determined by superimposing the two images and intersecting their major axes. Similarly, the core location is obtained by intersecting the image axes, starting from the locations of the two telescopes (assuming that all telescope dishes are in a common plane). Thus, the four parameters describing the major axes of the two images can be used to determine the four parameters describing the shower geometry - a direction in space and an impact point in a reference plane. It may be worth noting that the stereoscopic reconstruction of air showers makes the single (trivial) assumption than on average Cherenkov images are symmetric with respect to the (image of the) shower axis. \begin{figure}[htb] \begin{center} \mbox{\epsfxsize10.0cm \epsffile{image.eps}} \caption{Cherenkov image of a gamma-ray shower and its interpretation. The major axis of the image approximates the image of the shower axis; the image of the gamma-ray source is located on the image of the shower axis. Due to fluctuations in the shower development and in the imaging process, the center of gravity (c.o.g.) of the image can be displaced from the shower axis, and also the orientation of the image can deviate. These errors are indicated as an error ellipse for the image c.o.g., and an error on the image orientation. Taking into account these errors, the image of the source is constrained to the region between the dashed-dotted lines. Since with a simple elliptical parameterization there is a head-tail ambiguity of the image, the source can be located on either side of the image. The shape of the image, in particular its ellipticity, can be used to estimate the shower impact parameter relative to the telescope and hence the distance $d$ between the image of the source and the c.o.g. of the Cherenkov image.} \label{fig_image} \end{center} \end{figure} If a shower is observed by more than two telescopes, the shower geometry is overconstrained and some kind of suitable averaging or fitting procedure is required to extract optimum shower parameters from the information obtained from the different views. Particularly crucial is the case where the quality of the information provided by the different telescopes differs significantly, e.g. because one telescope is well within the light pool and sees a large intensity of Cherenkov light, whereas a distant telescope may see barely enough light to provide a meaningful image. Ideally, the reconstruction algorithm should take this difference in image quality into account. This paper reviews a number of different algorithms and describes tests of their performance based on the large sample of gamma rays \cite{hegra_mkn} collected with the HEGRA IACT system during the 1997 outburst of Mkn 501. \section{Techniques to reconstruct the shower geometry} To reconstruct shower parameters from the telescope images, two alternative approaches can be followed: \begin{description} \item[Using image parameters.] Images provided by the different telescopes are analyzed individually, and their key features are summarized in a small number of parameters (usually the well-known Hillas parameters). Shower parameters are derived on the basis of these image parameters. \item[Using the full image information.] A global optimization procure is applied to derive the shower parameters directly from the amplitudes measured in the individual pixels of all cameras. An example of such techniques are global fits, where parameterized shower images or image templates are matched to the images observed in the different telescopes \cite{hegra_fit,cat_fit}. \end{description} In this paper, we will concentrate primarily on methods of the first type. They are easier to implement, and usually require significantly less processing time. Often, analytical solutions for the shower parameters can be derived, and one does not have to worry about issues which arise in numerical optimization procedures, such as the choice of proper starting values and the convergence to the global optimum. For completeness and to serve as a reference, also results based a technique of the second type will be given. We will discuss seven different algorithms, six based on the Hillas image parameters and one based on a global fit to pixel amplitudes. \begin{description} \item[Algorithm 1.] For all pairs of telescopes, the image axes, derived using the Hillas parameterization, are intersected. In case of $N$ telescope images, the resulting $N(N-1)/2$ intersection points are averaged, weighted with the sine of the angle between the image axes, to take into account that image pairs with a large stereo angle provide the most precise determination of the shower axis. Similarly, the core location can be obtained by intersecting the image axes, starting from the telescope locations. This technique is used, e.g., for all published results from the HEGRA IACT system. It is illustrated in Fig.~\ref{fig_method}(a). \item[Algorithm 2.] A drawback of Algorithm 1 is that differences in the quality of the images in the different telescopes are not taken into account. The algorithm can be improved by determining the uncertainty in the determination of the image c.o.g. and in the direction of the image axis, and by taking the resulting errors (see Fig.~\ref{fig_image}) into account when intersecting the image axes, see Fig.~\ref{fig_method}(b). For $N$ intersecting lines with fixed error bands, the optimum solution can be derived analytically. Since the width of the error band associated with each image depends on the distance $d$ to the image c.o.g., one needs to iterate, but the result is stable after two iterations. This method also provides errors on the shower parameters. \item[Algorithm 3.] The image shape contains information on the distance $d$ (Fig.~\ref{fig_image}) between the image c.o.g. and the image of the source. In particular, the ratio of image {\em width} over image {\em length} can serve as a measure for $d$ (see also \cite{whipple_wl}). Smaller {\em width/length} implies large impact distance and large $d$. Together with a suitable error estimate for $d$, the location, orientation and shape of each image constrains the image of the source to two elliptical regions on both sides of the image (reflecting the left-right ambiguity inherent in the parameterization of shower images), see Fig.~\ref{fig_method}(c). For two or more images, these error ellipses can be combined analytically to yield the optimum shower direction and its errors. An analogous method determines the core location. \item[Algorithm 4.] Algorithms 1,2,3 determine independently the shower direction and the core location. Since the measurement of the image orientation is used both in the determination of the shower direction and of the shower core, a combined determination of core and direction should yield improved results. Technically, for a given shower geometry, the predicted image center lines are calculated, and a $\chi^2$ is defined measuring the agreement of the observed image and its orientation with this prediction. Shower geometry is chosen to minimize the sum of $\chi^2$ over telescope images. This method is illustrated in Fig.~\ref{fig_method}(d). \item[Algorithm 5.] Algorithm 4 can be augmented to include the estimate of $d$ from the {\em width/length}-ratio (see Algorithm 3), by adding corresponding terms to the $\chi^2$. \item[Algorithm 6.] Similar to algorithm 4, this algorithm -- proposed by Hillas \cite{global_width} -- calculates the image axes for a given shower geometry, and varies the shower parameters such as to minimize the sum of the squared distances of pixels to the axes, weighted with the pixel amplitudes. This technique is analogous to the determination of the image axis for single images, except that the weighted sum of pixel distances is minimized for the entire set of images together, rather than for single images \footnote{At first glance, Algorithm 6 may appear as an algorithm using the full pixel information rather than the image parameters. However, given the Hillas image parameters, equivalent to the moments of inertia of the image with respect to the major and minor axis, on can easily calculate the second moment with respect to an arbitrary axis.}. \item[Algorithm 7.] This last algorithms makes use of the full image information, by comparing the measured images with parameterizations of shower images, considering the shower geometry, the energy and the height of the shower maximum as free parameters which are chosen to minimize the $\chi^2$ describing the agreement between model and data. On the basis of Monte-Carlo simulations, the technique was discussed in \cite{hegra_fit}; a similar method was presented in \cite{cat_fit}. The method used here differs from \cite{hegra_fit} in a different choice of weights, which result in improved convergence. \end{description} In our implementation, Algorithms 1,2 and 3 use analytic expressions (with one iteration in case of Algorithm 3), whereas Algorithms 4 through 7 are based on numerical minimization procedures. \begin{figure}[htb] \begin{center} \mbox{\epsfxsize14.0cm \epsffile{method.eps}} \caption{Illustration of different techniques to determined the shower direction from multiple Cherenkov images. (a) Intersecting pairs of image axes, followed by an averaging over intersection points. (b) Intersecting image axes taking into account the errors on image location and image orientation, resulting in an error ellipse for the image of the source. (c) Using in addition the {\em width/length}-ratio to constrain the source image to two regions on either side of an image. (d) Optimizing the shower geometry such that the predicted image axes best match the observed images.} \label{fig_method} \end{center} \end{figure} \section{Data sample} To test the different algorithms and to experimentally determine the directional resolution achieved with each algorithm, data collected with the HEGRA IACT system during the 1997 outburst of Mkn 501 were used. In the HEGRA IACT system is located on the site of the Observatorio del Roque de las Muchachos on the Canary Island of La Palma at 2200 m asl. In 1997, the IACT system comprised four telescopes, located in the center and at three sides of a square with roughly 100~m side length. A fifth telescope at the remaining corner was integrated into the system in 1998. The telescopes are identical, equipped with 8.5~m$^2$ mirrors with 5~m focal length, and with 271-pixel cameras with a diameter of the field of view of $4.3^\circ$, and an equivalent pixel size of $0.25^\circ$. Detailed about the hardware and the data analysis can be found in \cite{hegra_perf,hegra_trigger,hegra_mkn,hermann_padua}. The trigger condition for individual telescopes requires a coincidence of two pixels above a threshold of 10 photoelectrons (before June `97) or 8 photoelectrons (after June `97). For typical gamma-ray images, cameras trigger once the image has more than about 40 photoelectrons. The HEGRA IACT system as a whole is triggered and data are recorded whenever at least two telescopes trigger in coincidence (see \cite{hegra_trigger} for details on the trigger system). In the design of the HEGRA cameras and their electronics, an important aspect was that one wanted to read out not only those telescopes which had triggered, but also the remaining telescopes, which will shower fainter, but frequently still usable images. Camera signals are digitized continuously by 120~MHz Flash-ADCs and are stored in a 34~$\mu$s ring buffer. A coincidence trigger of at least two telescopes is generated with a delay of 1 to 2 $\mu$s; After such a trigger, the readout system addresses the relevant locations in the Flash-ADC memory and extracts the signals. In 1997, Mkn 501 was observed in the so-called wobble mode, with the source offset by $0.5^\circ$ in declination from the optical axis of the telescopes. The offset alternated every 20~min. A region offset by the same amount, but in the opposite direction, is used as a control region and for background subtraction. The analysis is based on data taken during three new-moon periods, where the gamma-ray flux from Mkn 501 was particularly high. Only data at small zenith angles, below $20^\circ$, are included; these showers behave essentially like ``ideal'' vertical showers, at least as far as the angular reconstruction is concerned. The usual selections concerning data quality were applied, see \cite{hegra_mkn}. In total, the sample comprises 100748 events within $0.5^\circ$ from the source, and 80878 events in the equivalent off-source region. Images were flat-fielded, and corrections for pointing errors of the telescopes were applied \cite{hegra_pointing}. Image parameters were determined by selecting ``image pixels'' as those pixels which either have a signal of 6 or more photoelectrons, or which have a signal of at least 3 photoelectrons and are adjacent to a pixel with 6 or more photoelectrons. \section{Errors assigned to image parameters, and angular resolution} Some of the algorithms discussed above require errors on the image parameters as input for the reconstruction of the shower axis. The relevant image parameters are the coordinates $(x,y)$ of the center of gravity of the images, and the direction $\theta$ of the major axis of the image. To parameterize the errors on the center of gravity, it is more convenient to use a coordinate system where axes $(u,v)$ are defined by the major ($u$) and minor ($v$) axes of the image. In this system, the errors on $u$, $v$, and on the orientation $\theta$ of the image should be essentially uncorrelated. The errors (in units of degr.) were parameterized on the basis of Monte-Carlo simulations \cite{hegra_mc}: $$ \Delta v = \left\{ {0.03 \over A} + 0.009^2 \right\}^{1 \over 2} f(w) ~~~~,~~~~~~~ f(w) = \left\{ \begin{array}{cl} 1 & \mbox{~~if~} w < 0.08 \\ w/0.08 & \mbox{~~if~} w \ge 0.08 \end{array} \right. $$ and $$ \Delta \theta = \left\{ {\left( {600 \over A}\right) }^{1.5} + 1.1^2 \right\} ^{1 \over 2} +45 \left( {w \over l} - 0.2 \right)^2 $$ Here, $A$ denotes the number of photoelectrons in the image ({\em size}), $w$ the image {\em width} and $l$ the {\em length}. The error in $u$ is not very relevant as long as the major axis of the image points more or less towards the source; $\Delta u = 2 \Delta v$ was used. As a first check to see if these errors describe the uncertainties in the real data, the distribution in {\em miss} was plotted for individual telescopes of the (background subtracted) Mkn 501 gamma-ray sample, with the (signed) {\em miss} parameter normalized to the expected error. The {\em miss} parameter describes the distance between the image axis and the point on the camera which corresponds to the image of the source. The error on {\em miss} is $\Delta^2_{miss} = d^2 \Delta \theta^2 + \Delta v^2$. For typical values of $d \approx 1^\circ$ the $\Delta \theta$ term gives the dominant contribution. The normalized {\em miss} distribution has an rms width of 1.06, and its central part is well described by a Gaussian with a width of 0.90, indicating both that the errors estimated are accurate within 10\%, and that alignment errors of the telescopes are small on the scale of the resolution. \begin{figure} \begin{center} \mbox{\epsfxsize7.0cm \epsffile{ctsh_q.eps}} \caption{(a) Predicted uncertainty in the measurement of the (projected) direction of the shower axis, for events with at least two triggered telescopes (full lines) and for events where all four telescopes triggered (dashed). Only triggered telescopes are used in the reconstruction. (b) Deviation between the measured shower axis and the direction to Mkn 501 for events with a predicted angular error of less than $0.04^\circ$, after statistical subtraction of the background. The curve represents a Gaussian fit with a width of $0.037^\circ$. (c) Experimental angular resolution (in projections), determined using a Gaussian fit, as a function of the predicted error of the measurement.} \label{fig_sel} \end{center} \end{figure} To verify that also errors on the shower direction can be reliably calculated by propagating the errors on the image parameters, events were reconstructed with Algorithm 2 and those events were selected where the predicted error on the shower direction was less than $0.04^\circ$ (Fig.~\ref{fig_sel}(a)). For these events, the distribution in the difference between the reconstructed shower direction and the source, Mkn 501, was plotted (Fig.~\ref{fig_sel}(b)), projected onto two orthogonal axes. Indeed, for this subsample of events, a (projected) angular resolution of $0.037^\circ$ is obtained, confirming the validity of the approach to estimate the errors, by treating the image parameters obtained by the different telescopes as independent measurements. Here and in the following, `angular resolution' refers to the width of the angular distribution of shower axes in a projection. If angular resolution is defined as the half opening angle of a cone in space, which contains 68\% of the events, the numerical values are a factor 1.5 larger (assuming a Gaussian distribution in the errors). Fig.~\ref{fig_sel}(c) finally shows the measured angular resolution as a function of the predicted resolution, demonstrating good agreement except for the tail of events with very large predicted errors ($> 0.2^\circ$), where the measured angular resolution is slightly better than expected (most likely due to imperfections in the parameterizations of the errors). Hence, already with the extremely simple and fast Algorithm 2 one can reliably reject events with poorly reconstructed showers, which is important, e.g., for the determination of energy spectra. In addition to errors on the image parameters, Algorithms 3 and 5 require an estimate of $d$ based on the image shape; we used the empirical relations $$ d = 1.4 - 1.25 {w \over l} ~~~~~~,~~~~~~ \Delta d = \max \left( {2.5 \over \sqrt{A}}~,~0.15 \right) $$ \section{Comparison of reconstruction techniques} \begin{figure} \begin{center} \mbox{\epsfxsize10.0cm \epsffile{ctsh_final3_log.epsi}} \caption{Angular resolution obtained from the Mkn 501 data with the various algorithms, for different data samples and reconstruction modes. Full squares: events with exactly two triggered telescopes, using only these two telescopes; open circles: events with exactly two triggered telescopes, using images in all four telescopes; full circles: events where all four telescopes triggered and all images are used.} \label{fig_resol} \end{center} \end{figure} The angular resolutions obtained with the seven algorithms described above are summarized in Fig.~\ref{fig_resol}, for three characteristic data samples chosen to emphasize the specific features on the algorithms: \begin{description} \item[2-Telescope events] (full squares). In these events, exactly two telescopes have triggered, and only these two telescopes are used for the reconstruction. The 2-Telescope sample serves primarily to verify that all algorithms work properly; unless additional shape information is used (such as in Algorithms 3, 5, and 7), all algorithms should give identical results if only two images are used in the reconstruction \footnote{It should be noted that there is a big difference between the 2-Telescope sample, where exactly two of the four telescopes triggered, and samples (``2/x'') where two telescopes are used for the reconstruction, regardless of the state of the other two. The 2-Telescope sample selects events which either have energies near the trigger threshold, or which have quite distant cores. A 2/x-Telescope sample yields for Algorithm 2 a resolution of about $0.10^\circ$, compared to the $0.14^\circ$ for the 2-Telescope sample -- see below.}. \item[2+2-Telescope events] (open circles). In these events, exactly two telescopes have triggered, but images in the other two untriggered telescopes are included in the reconstruction. These events represent a particular challenge to reconstruction algorithms, since they combine images of very different quality. Triggered images contain a mean number of about 150 photoelectrons, compared to about 30 photoelectrons in images which did not trigger. \item[4-Telescope events] (full circles). In these events, all four telescopes have triggered and are used in the reconstruction. This class of events will obviously provide the best angular resolution. \end{description} For the 2-Telescope sample (full squares in Fig.~\ref{fig_resol}), Algorithms 1, 2, 4 and 6 do indeed provide the identical angular resolution. Algorithms 3, 5 and 7 -- which add shape information -- give significantly improved resolution. This improvement can be traced to events with small stereo angles, i.e. with shower cores along the line connecting the two telescopes; for such events, the purely geometrical reconstruction fails and the otherwise relatively poor shape information helps to stabilize the reconstruction. Adding now in the reconstruction the faint images of the other two telescopes, which did not trigger -- the 2+2-Telescope sample (open circles) -- one mixes images of rather different quality. If all images are combined with equal weight, as in Algorithm 1, the faint images hurt the resolution; the resulting resolution is worse than if only the two triggered telescopes are used. In all other algorithms, the faint images weigh less than the good images, either because explicitly larger errors on the image parameters are assigned (Algorithms 2-5), or because the effect of the images is weighted with the number of photoelectrons they contain (Algorithms 6,7). These algorithms improve the angular resolution by about 20\% to 30\% compared to the 2-Telescope sample. Little is gained by adding the shape information via the $d(w/l)$ relation; with four views there is always at least one reasonably large stereo angle. In many respects, the 4-Telescope sample (full circles) is less critical than the 2+2-Telescope sample, since the differences between the quality of the four images are not nearly as big. Hence, it is no surprise that the variation between algorithms is smaller, ranging from a resolution of $0.072^\circ$ for the worst case (Algorithm 1) to $0.056^\circ$ for the best case (Algorithm 7). \begin{figure} \begin{center} \mbox{\epsfxsize14.0cm \epsffile{dircore.eps}} \caption{(a) Reconstruction of the shower direction, by intersecting the image axes, starting from the image c.o.g. (b) Reconstruction of the shower core by intersecting the image axes, starting from the telescope locations. (c) $\chi^2$ describing the consistency in the determination of shower cores in overconstrained events, vs. $\chi^2$ describing the consistency of the determination of the shower direction, for background-subtracted gamma-ray events from Mkn 501.} \label{fig_dircore} \end{center} \end{figure} One may wonder why the joint fits of the shower direction and of the core location (Algorithms 4, 5, 6) do not provide significant improvements. The explanation is relatively simple, and is illustrated in Fig.~\ref{fig_dircore}. The geometrical figure describing the determination of the direction (Fig.~\ref{fig_dircore}(a)) and the figure describing the determination of the core (Fig.~\ref{fig_dircore}(b)) are essentially scaled versions of each other, since the {\em distance} parameter $d$ of the image is approximately proportional to the distance $r$ from the telescope to the core location. The main difference is that in the determination of the shower direction, the error on the position of the image c.o.g. enters in addition to the error on the image orientation; in the core determination, only the latter matters. Since the error on the c.o.g. is usually of little relevance compared to the error on the orientation, the joint fit does not add additional constraints. Indeed, one finds that in Algorithm 2, the $\chi^2$ describing how well the telescopes match in the determination of the shower direction, and the $\chi^2$ of the core determination are highly correlated (Fig.~\ref{fig_dircore}(c)). Fitting the full image information (Algorithm 7) results in only very modest improvements compared to the simpler Algorithms 2 - 6. At least with the pixel size of $0.25^\circ$ of the HEGRA cameras, the Hillas image parameters seem to very efficiently capture the essence of the information contained in the images. \section{Dependence of the angular resolution on the number of telescopes used in the reconstruction} An interesting question is how the angular resolution depends on the number of telescopes $N_{tel}$ used in the reconstruction. If the individual images can be considered as independent, the resolution should improve like $1/\sqrt{N_{tel}}$. However, at some point shower fluctuations will start to dominate the resolution. As mentioned above, to address this issue one cannot simply use the event samples where exactly 2, 3 or 4 telescopes have triggered, since the 2-telescope sample is biased towards low-energy or distant showers, whereas in the 4-telescope sample central high-energy events are enhanced. To start from identical event samples and to avoid a ``trigger bias'', the investigation was based on the 4-telescope sample, but only a subset of telescopes was used to reconstruct the shower. The resulting resolutions are illustrated in Fig.~\ref{fig_resol}(a), for Algorithms 2 and 3. (Note that Algorithm 3 can reconstruct the shower direction from a single image, apart from the head-tail ambiguity.) Except for the minimum number of telescopes -- 1 for Algorithm 3 and 2 for Algorithm 2 -- data are consistent with a $1/\sqrt{N_{tel}}$-dependence. The issue was further was explored on the basis of Monte-Carlo simulations for the HESS telescope system \cite{hess_mc}. These simulations used an array of 589 telescopes, arranged as a square grid of 31 x 19 telescopes, spaced 33.3~m. In the analysis, arbitrary subsets of telescopes can be selected. The sets studied here include a set with all telescopes turned on, a set where every other telescope is active, and sets with telescopes on square grids with an effective spacing of 67~m, 100~m, 133~m, and 167~m. Only showers well contained within the array were considered. Fig.~\ref{fig_resol1}(b) shows the resulting angular resolution as a function of the mean number of telescopes used in the reconstruction. The $1/\sqrt{N_{tel}}$-dependence of the resolution holds up to about 50 telescopes used per event, and resolutions better than $0.03^\circ$. For even higher telescope numbers, the dependence appears to flatten somewhat. \begin{figure} \begin{center} \mbox{\epsfxsize8.0cm \epsffile{ctshres.eps}} \caption{(a) Angular resolution obtained from the Mkn 501 data with Algorithm 2 (full circles) and Algorithm 3 (open circles), for events where all four telescopes triggered, but only a (random) subset of telescopes is used in the reconstruction. The curve illustrates a $1/\sqrt{N_{tel}}$-dependence. (b) Angular resolution obtained in Monte-Carlo studies using an array of telescopes, as a function of the average number of telescopes used in the reconstruction. Shower energies range from 0.5 TeV to 1 TeV. The curve illustrates a $1/\sqrt{N_{tel}}$-dependence. (The telescope characteristics differ from those of the HEGRA telescopes, and the resolutions cannot be compared directly.)} \label{fig_resol1} \end{center} \end{figure} \section{Concluding remarks} The main conclusions from these studies of different algorithms for the stereoscopic reconstruction of multi-telescope IACT events are: \begin{itemize} \item By assigning and properly propagating errors of the image parameters, reliable error estimates for the shower direction can be obtained. It is possible to select subsamples with improved angular resolution -- less than $0.05^\circ$, e.g. -- for special purposes, such as to study the size of the source, or to exclude poorly reconstructed events. \item In particular when combining multiple and partly redundant images of rather different quality, the reconstruction algorithm must properly account for these differences. \item Using image shape information to constrain the direction of the shower axis helps in the case of 2-telescope events with small stereo angles; for events with more than two telescopes, the improvement is small. \item Compared to the simple, robust and fast Algorithms 2 and 3, the more `fancy' Algorithms 4, 5 and 6 as well as the rather sophisticated image fitting procedure of algorithm 7 give only modest improvements. For most practical purposes; Algorithms 2 or 3 may represent the simplest and best choice. \end{itemize} Of course, these conclusions hold primarily for the HEGRA Cherenkov telescopes. To which extent they can be applied to other IACT systems depends on the degree of similarity in the trigger concept and the layout of the cameras. \section*{Acknowledgements} The support of the HEGRA experiment by the German Ministry for Research and Technology BMBF and by the Spanish Research Council CYCIT is acknowledged. We are grateful to the Instituto de Astrofisica de Canarias for the use of the site and for providing excellent working conditions. We thank the other members of the HEGRA CT group, who participated in the construction, installation, and operation of the telescopes. We gratefully acknowledge the technical support staff of Heidelberg, Kiel, Munich, and Yerevan.
\section{Introduction} Polyakov's proposal~\cite{Polyakov:a} of using type 0A/0B string theory for describing non-supersymmetric Yang-Mills theories is a promising and novel attempt at formulating the QCD string and provides an alternative way of extending the AdS/CFT correspondence~\cite{Maldacena, Gubser:a, Witten:a} beyond the realm of supersymmetric field theories. Simply put, the proposal postulates using a closed string theory in $d\leq 10$ space-time dimensions with world-sheet supersymmetry and a diagonal GSO projection that removes all space-time fermions, thus yielding a non-supersymmetric theory in target space. One can refer to such theories for which $d<10$ as ``non-critical'' although the name can be confusing because Weyl invariance is recovered once the conformal factor of the world-sheet metric is counted among the other space-time coordinates. Throughout this paper, $d$ will always include such Liouville mode and hence Weyl invariance is retained, at least at the lowest level in the sigma-model expansion. So far, most of the literature following~\cite{Polyakov:a} has analyzed the most conservative case of non-supersymmetric theories in $d=10$~\cite{Klebanov:a, Klebanov:b, Klebanov:c, Minahan:a, Minahan:b, Alishahiha}. Such theories~\cite{Dixon, Seiberg} are often referred to as type 0A or 0B, depending on the particular GSO projection employed. Their open string descendants were analyzed in~\cite{Bianchi, Sagnotti, Angelantonj, Bergman}. However, it is our opinion that the ``non-critical'' scenario~\cite{Alvarez:a, Ferretti:a, Alvarez:b, Costa, Armoni, Zhou} must be taken seriously and will give rise to additional interesting models that are not accessible at $d=10$. Throughout the paper we shall continue to refer to these lower dimensional theories as type 0A or 0B in even dimensions, depending on the choice of chirality in the GSO projection. In odd space-time dimensions there is only one such theory due to the lack of chirality. We shall refer to it as type 0AB. Although the full conformal field theory corresponding to the above non-critical string theory has not yet been constructed, there are some indications that such a construction is indeed possible: Consider the issue of modular invariance. Define the fermionic traces for the pair of Majorana--Weyl fermions $\psi(z)$, $\bar\psi(z)$ as\footnote{We use the notation of~\cite{Polchinski:a}. Here $q=\exp(2\pi i \tau)$.} \begin{eqnarray} Z^0_0(\tau) &=& {\mathrm tr\;}_{\mathrm NS} \bigg(q^N\bigg) \nonumber \\ Z^0_1(\tau) &=& {\mathrm tr\;}_{\mathrm NS} \bigg((-)^F q^N\bigg) \nonumber \\ Z^1_0(\tau) &=& {\mathrm tr\;}_{\mathrm R} \bigg(q^N\bigg) \nonumber \\ Z^1_1(\tau) &=& {\mathrm tr\;}_{\mathrm R} \bigg((-)^F q^N\bigg). \label{zetas} \end{eqnarray} In the ordinary type II string theory in $d$ dimensions, modular invariance requires that the holomorphic contribution of the fermions to the partition function \begin{equation} Z^0_0(\tau)^{(d-2)/2} - Z^0_1(\tau)^{(d-2)/2} - Z^1_0(\tau)^{(d-2)/2} \pm Z^1_1(\tau)^{(d-2)/2} \end{equation} and the analogous antiholomorphic contribution be separately modular invariant up to an overall opposite phase. It is well known that the lowest dimension where the product of the holomorphic contribution and the antiholomorphic contribution is modular invariant is $d=10$. In type 0 theories, the joined contributions of the holomorphic and the antiholomorphic sectors give rise to \begin{equation} |Z^0_0(\tau)|^{d-2} + |Z^0_1(\tau)|^{d-2} + |Z^1_0(\tau)|^{d-2} \pm |Z^1_1(\tau)|^{d-2}, \end{equation} which is modular invariant for any $d$. Of course, for $d\not=10$ the explicit expressions for the fermionic traces will be modified because of the changes in the spectrum, but we view the above as an indication that the continuation ``off-criticality'' is more likely to work for the type 0 string than for the usual type II. Another objection that needs to be addressed is the ``$d=2$ barrier''. Let us briefly recall the physics behind this problem as presented e.g.~in~\cite{Polchinski:a}. Because we do not yet know how to describe RR fields at the level of the sigma-model we are forced to discuss this argument in the case of the bosonic theory. In this context it is well know that there exists an exact CFT solution in any $d$ in which the only non-zero background fields are a flat metric $g_{MN}$, a linearly rising dilaton $\Phi$ and an exponential tachyon $T$ \begin{equation} g_{MN}=\eta_{MN}, \quad\Phi=\sqrt{\frac{26-d}{6}}X^1, \quad T=\exp\left(\left( \sqrt{\frac{26-d}{6}}-\ \sqrt{\frac{2-d}{6}}\right)X^1 \right). \end{equation} For $d\leq 2$ the background tachyon is exponentially rising, preventing the string from entering the region of strong coupling. Moreover, fluctuations around the tachyon background are stable and thus the theory is well defined. On the contrary, for $d>2$, the background tachyon oscillates. In principle this could still act as a cutoff for the string coupling but the fluctuations have some negative frequency square modes and the theory becomes unstable. As stressed in~\cite{Polyakov:a, Polyakov:b} we should not think of the ``$d=2$ barrier'' as a no-go theorem but rather as an indication that solutions for $d>2$ will necessarily involve a curved space-time metric. This is the type of situation that is of interest in the connection with gauge theory so, in a sense, it is to be expected that flat space-time be ruled out. Unfortunately, we do not yet have an example of an exact CFT of this type and we are forced to work order by order in $\alpha^\prime$ at the level of the effective action. But it should be clear that there are no a priori reasons for why there should not exist an exact solution. A third encouraging sign comes from the analysis of the RR sector performed in Section 3. By making some plausible assumptions about the massless degrees of freedom it is possible to construct a rather compelling picture of the RR sectors in various dimensions and their couplings, including Chern--Simons terms. For instance, the necessity of doubling the RR spectrum in $d=4$ or $d=8$ Minkowski space-time is seen as coming from the fact that there are no real self-dual forms in these dimensions. Finally, let us note that considering $d<10$ from the sigma model point of view is a very natural thing to do if the string theory has a perturbative tachyon. The only way for a theory with a perturbative tachyon to make sense is if there exists a mechanism through which the tachyon field condenses by acquiring a vacuum expectation value. The tachyon potential at that point will then give rise to a tree-level contribution to the cosmological constant by shifting the central charge. Since the effective central charge is going to be different from zero anyway\footnote{It seems unnatural and there is no symmetry argument for which the tachyon potential should vanish at that point.}, one is led to consider the theory with the most general value for the effective central charge \begin{equation} c_{\rm eff.} = 10 - d - \frac{1}{2}V\bigg(\langle T \rangle\bigg). \end{equation} {}From the target space point of view this acts as a contribution to the cosmological constant and thus shows that, for $d>2$, one should look at curved space-times. None of the above points constitute a proof that conformally invariant solutions to the type 0A/B string exist for arbitrary $d$ but we view them as strong indications that such construction is possible. Having taken this as our basic assumption, throughout the paper we work at the level of the effective action to one loop in $\alpha^\prime$, i.e. the gravity level without higher order corrections. The form of the action of the type 0 gravity, can be determined perturbatively, with a certain number of ambiguities involving the tachyon potential and the coupling to the RR fields, which are essentially there because of lack of supersymmetry on the target space. However, supersymmetry on the world sheet allows one to make some statements on these terms, as we will see in Section 3. We will stay generic and will be able to show that in any dimension there exists a set of exact solutions of the classical equations of motion, which give AdS metric and involve a non-zero RR field, other than constant dilaton and tachyon. Such solutions depend only on a {\it finite} number of parameters for which a string-theoretical derivation is still lacking. The solutions found represent Polyakov's conformal fixed points in the dual gauge theory --- they support a condensed tachyon, but the first issue to be addressed is the stability against quantum fluctuations of the fields. Because of the mixing of several of these fields it is not enough to analyze the tachyon itself, one has to disentangle the full set of fluctuations~\cite{Ferretti:b}. For simplicity, we restrict to the case where the space is $d$-dimensional AdS, so that there are no KK modes to be worried about. There the analysis simplifies considerably, but it is still non trivial because of dilaton-tachyon mixing. Another issue that was raised in~\cite{Polyakov:a} is the field theory interpretation of these solutions. There it was claimed that they may represent an interacting UV conformal point. To address this issue one needs to show that these gravity solutions represent in fact a point in wider space, that is actually the RG phase diagram of the field theories we have at hand. In this enlarged theory space there may be more than one conformal solution and there exist trajectories which interpolate between these points. These considerations have already been realized in the framework of type IIB supergravity, where the RG flow is believed to be driven by operators dual to scalar fields coming from KK reduction in the compact directions~\cite{Distler, Girardello:a, Girardello:b, Porrati}, or by the running of the dilaton \cite{Kehagias:a, Gubser:b, deMelloKoch, Kehagias:b}. It turns out that type 0 gravity provides an example of this general feature of the AdS/CFT correspondence, as well. In fact, the physics is already captured by the set of solutions involving no compact space. Even without including KK modes, the tachyon will mix with the dilaton field, and generate on the field theory side a RG flow that connects interacting conformal fixed points. \section{A look at the type 0A/B theory in $d=10$} In this section we briefly summarize some known facts about the perturbative properties of these theories. We will thus set $d=10$ throughout this section and consider perturbation theory around the (unstable) vacuum. We will, however, point out the various places where modifications occur when one is considering $d\not= 10$. We will return to these changes in Section 3 where we present the detailed structure of the RR terms in $d\leq10$. In the notation of~\cite{Polchinski:a}, there are (up to equivalences) only four consistent ways of combining the various sectors of the ${\cal N}=(1,1)$ NSR closed oriented string in $d=10$: \begin{eqnarray} {\mathbf IIA}\;\;&:&\;\; (NS+,NS+)\oplus(NS+,R+)\oplus(R-,NS+)\oplus(R+,R-)\nonumber \\ {\mathbf IIB}\;\;&:&\;\; (NS+,NS+)\oplus(NS+,R+)\oplus(R+,NS+)\oplus(R+,R+)\nonumber \\ {\mathbf 0A}\;\;&:&\;\; (NS+,NS+)\oplus(NS-,NS-)\oplus(R+,R-)\oplus(R-,R+)\nonumber \\ {\mathbf 0B}\;\;&:&\;\; (NS+,NS+)\oplus(NS-,NS-)\oplus(R+,R+)\oplus(R-,R-). \label{GSO} \end{eqnarray} The first two are the usual type IIA/B superstring and the last two are those of interest here. The massless fields of the theory are the dilaton $\Phi$, graviton $g_{MN}$ and two-form $B_{MN}$ coming from the $(NS+,NS+)$ sector and twice as many RR fields coming from the doubled RR sectors. There is a tachyon $T$ from the $(NS-,NS-)$ and no fermionic mode at all. All other modes are massive. \subsection{Selection rules for $d=10$} The lack of space-time fermions means that the theory has no space-time supersymmetry; however, the presence of ${\cal N}=(1,1)$ world-sheet supersymmetry has the following simplifying features~\cite{Klebanov:a}: \begin{itemize} \item[a)] All tree-level correlators involving an odd number of tachyons and only $(NS+, NS+)$ vertex operators are zero. This can be seen from the explicit form of the vertex operators. The vertex operator for the massless $(NS+, NS+)$ states and for the tachyon in the $(NS-, NS-)$ sector are, again in the notation of~\cite{Polchinski:a} \begin{eqnarray} V^{0,0}_{(NS+, NS+)}&=&-(i\partial X^M +k\cdot\psi~ \psi^M) (i\bar\partial X^N +k\cdot\tilde\psi~ \tilde\psi^N)~ {\mathrm e}^{ik\cdot X}\nonumber\\ V^{-1,-1}_{(NS+, NS+)}&=&{\mathrm e}^{-\phi-\tilde\phi}~ \psi^M\tilde\psi^N~ {\mathrm e}^{ik\cdot X}\nonumber\\ V^{0,0}_{(NS-, NS-)}&=&k\cdot\psi~ k\cdot\tilde\psi~ {\mathrm e}^{ik\cdot X}\nonumber\\ V^{-1,-1}_{(NS-, NS-)}&=&{\mathrm e}^{-\phi-\tilde\phi}~ {\mathrm e}^{ik\cdot X}.\label{nsvertex} \end{eqnarray} Since on the sphere we need to take any two of the above in the $(-1,-1)$ picture and all the rest in the $(0,0)$ picture it is clear that we will always end up with the correlation function of an odd number of $\psi$'s which vanishes. In particular, notice that in this vacuum $\langle T \rangle = 0$ whereas we expect a tachyon condensate in the true vacuum. \item[b)] All tree-level correlators involving an odd number of tachyons, any number of fields from the $(NS+, NS+)$ and two RR fields from the same sector must vanish. For instance, if the amplitude with one tachyon did not vanish there would be a tachyon pole in some tree-level correlation function of the corresponding type II theory. The vanishing of the amplitude can also be seen at the level of the correlation functions of the vertex operators. We can always use the vertex operators for the RR fields in the $(-1/2, -1/2)$ picture -- written as a bispinor it reads \begin{equation} V^{-1/2,-1/2}_{(R\pm, R\pm)}={\mathrm e}^{-\phi/2-\tilde\phi/2}~ \Theta_\alpha\tilde\Theta_\beta~ {\mathrm e}^{ik\cdot X}~. \label{rrvertex} \end{equation} In a correlator that involves two such vertex operators, one remaining $(NS\pm,NS\pm)$ vertex (\ref{nsvertex}) needs to be taken in the $(-1,-1)$ picture yielding an overall even number of $\psi$'s. Schematically, this part of the correlator is proportional to $C\Gamma^{2n}$ and only connects spinors of opposite chirality because the charge conjugation matrix $C$ anticommutes with the chirality matrix $\Gamma_\chi$. This is also true in $d=6$ whereas in $d=4$ and $d=8$ the opposite is true, since now the matrix $C$ commutes with $\Gamma_\chi$. Hence, in $d=4$ and $d=8$ we expect to find a coupling between an odd number of tachyons and RR fields from the same sector and none between an odd number of tachyons and RR fields from the opposite sector. \item[c)] The reverse statement holds for an even number of tachyons (in particular no tachyon at all). The correlators will now involve an odd number of $\psi$'s and these vanish in $d=10$ between spinors of opposite chirality. Again, it is natural to expect that such a statement will hold in $d=6$ and be reversed in $d=4$ and $d=8$. \end{itemize} For odd space-time dimensions, both statements b) and c) are empty due to the lack of chirality and we do not expect any particular symmetry of the couplings. Statement c) is also consistent with the fact that in $d=10$ all correlation functions between $(NS+, NS+)$ fields and fields from \emph{one} given RR sector are the same as in type II theory, which is obvious because the vertex operators and their correlation functions are precisely the same and there is no loop in which other modes could propagate. In particular, it indicates that there are Chern--Simons terms in the effective action and that the various field strengths need to be modified by shifting them with a $B_{MN}$ dependent transformation just as in type II supergravity. It seems natural to postulate that such terms will be present also in lower dimensions in the cases where they are allowed by the symmetries of the problem. These simple properties and some explicit tree-level computations~\cite{Klebanov:a} allow one to write down, for the critical case, to first non-trivial order in $\alpha^\prime$, an effective action up to terms quadratic in the gauge fields. Below we shall present our proposed generalization of this action in any dimension including Chern--Simons terms. \subsection{D-branes in the $d=10$ theory} We conclude this section with a look at the type of D-branes in this theory~\cite{Bergman, Billo}. In this subsection we always work in $d=10$. Due to the doubling of the RR fields, there are twice as many D-branes as in the corresponding type II theory. Let us denote by $F$ and $F^\prime$ the RR fields from the two RR sectors of (\ref{GSO}) respectively\footnote{We refrain from using the notation $F$ and $\bar F$ used in~\cite{Klebanov:a} because we reserve such notation for the $d=4$ and $d=8$ case where such fields are complex conjugate of each other.}. For a given $p$ we have thus four types of elementary branes (counting the anti-branes) with charges $(Q=1, Q^\prime=1)$, $(Q=1, Q^\prime=-1)$, $(Q=-1, Q^\prime=1)$, $(Q=-1, Q^\prime=-1)$. Branes charged only with respect to, say, $F$ are possible but carry charge $(Q=2n, Q^\prime=0)$ in these units and thus can be built from the four constituents above. A quick way to understand the spectrum of massless excitations living on the world-volume of a stack of such branes is to consider the closed string exchange between two such branes, perform a modular transformation and read off the spectrum from the open string sector. Let us denote the usual open string traces in the various sectors as\footnote{We use the symbol ${\mathrm Tr\;}$ to denote the sum over all eight transverse bosonic and fermionic components, to distinguish it from the trace ${\mathrm tr\;}$ in (\ref{zetas}).} \begin{eqnarray} {\mathrm Tr\;}_{\mathrm NS} \bigg(q^N\bigg) &=& \left(\frac{f_3(q)}{f_1(q)} \right)^8\nonumber \\ {\mathrm Tr\;}_{\mathrm NS} \bigg((-)^F q^N\bigg) &=& \left(\frac{f_4(q)}{f_1(q)} \right)^8 \nonumber \\ {\mathrm Tr\;}_{\mathrm R} \bigg(q^N\bigg) &=& \left(\frac{f_2(q)}{f_1(q)} \right)^8, \end{eqnarray} where, as usual, \begin{eqnarray} f_1(q) &=& q^{1/12}\Pi(1-q^{2n}) \nonumber \\ f_2(q) &=& \sqrt{2} q^{1/12}\Pi(1+q^{2n}) \nonumber \\ f_3(q) &=& q^{-1/24}\Pi(1+q^{2n-1}) \nonumber \\ f_4(q) &=& q^{-1/24}\Pi(1-q^{2n-1}). \end{eqnarray} The closed string exchange can be written in terms of these functions by making a modular transformation $\tilde q={\mathrm e}^{-\pi\tilde t}\to q={\mathrm e}^{-\pi /\tilde t} = {\mathrm e}^{-\pi t}$. Let us denote by $H$ the light-cone oscillator part of the closed string Hamiltonian, and use the boundary states $|B\rangle$ to denote the branes (cf.~\cite{Hussain, DiVecchia} for a complete discussion of the boundary state formalism). Let us also introduce the shorthand notation \begin{equation} \left[NS+, NS+\right] =\langle B|{\mathrm e}^{-\tilde t H}|B\rangle_{(NS+, NS+)}, \end{equation} and similar expressions for the other sectors. Then, after a modular transformation: \begin{eqnarray} \left[NS+, NS+\right]&\to& \frac{1}{2} \left({\mathrm Tr\;}_{\mathrm NS} \bigg(q^N\bigg) - {\mathrm Tr\;}_{\mathrm R} \bigg(q^N\bigg) \right) = \frac{1}{2} \left(\frac{f_3(q)^8 - f_2(q)^8}{f_1(q)^8} \right) \nonumber\\ \left[NS-, NS-\right] &\to& \frac{1}{2} \left({\mathrm Tr\;}_{\mathrm NS} \bigg(q^N\bigg) + {\mathrm Tr\;}_{\mathrm R} \bigg(q^N\bigg) \right) = \frac{1}{2} \left(\frac{f_3(q)^8 + f_2(q)^8}{f_1(q)^8} \right) \nonumber\\ \left[R\pm, R\pm\right] &\to& -\frac{1}{2} {\mathrm Tr\;}_{\mathrm NS} \bigg((-)^F q^N\bigg) = -\frac{1}{2} \left(\frac{f_4(q)^8}{f_1(q)^8} \right).\\ \end{eqnarray} All RR sectors give the same contribution. There are three elementary cases to consider. First, consider the case of two like-like charged branes $(Q_1=1, Q^\prime_1=1)$ and $(Q_2=1, Q^\prime_2=1)$. Let us consider the type 0B case for definitiveness; all considerations apply to the type 0A case as well. In this case we have the following situation: \begin{equation} \left[NS+, NS+\right]+\left[NS-, NS-\right]+ \left[R+, R+\right]+ \left[R-, R-\right] \to \frac{f_3(q)^8 - f_4(q)^8}{f_1(q)^8}. \label{electric} \end{equation} {}From (\ref{electric}) we read off, with the wisdom of~\cite{Polchinski:b, Witten:b}, that the world-sheet theory has no tachyon, no fermions, and the same massless bosons as pure $d=10$ Yang-Mills theory dimensionally reduced to $p+1$ dimensions. Second, consider the case $(Q_1=1, Q^\prime_1=1)$ and $(Q_2=1, Q^\prime_2=-1)$. Here the exchange is \begin{equation} \left[NS+, NS+\right]-\left[NS-, NS-\right]+ \left[R+, R+\right]- \left[R-, R-\right] \to -\frac{f_2(q)^8}{f_1(q)^8}. \label{fermi} \end{equation} Perhaps the only subtlety is the minus sign in front of the $(NS-, NS-)$ exchange. This sign is fixed by looking at the coupling of the tachyon to the branes and noticing~\cite{Klebanov:a} that it has a coupling constant proportional to the product of the charges of the brane, thus yielding $Q_1 Q^\prime_1 \times Q_2 Q^\prime_2 = -1$. {}From (\ref{fermi}) we see that this brane configuration has only fermions on the world-volume. These two configurations were studied in~\cite{Klebanov:c}. Finally the brane/anti-brane case is given by $(Q_1=1, Q^\prime_1=1)$ and $(Q_2=-1, Q^\prime_2=-1)$ and corresponds to \begin{equation} \left[NS+, NS+\right]+\left[NS-, NS-\right]- \left[R+, R+\right]- \left[R-, R-\right] \to \frac{f_3(q)^8+f_4(q)^8}{f_1(q)^8}. \end{equation} There is an open string tachyon in this theory, just as in~\cite{Green, Banks}, signaling an instability of the theory. \section{Type 0 effective actions, Ramond--Ramond fields and Chern--Simons terms} In this section we present our proposal for the effective action of type 0 string theory in any dimension including the Chern--Simons couplings. Lacking a formulation from ``prime principles'', the identification of the RR sectors and their couplings requires a certain amount of guesswork. The picture that emerges, however, is quite simple and satisfying. We shall see, for instance, that it gives support to the idea that the RR sectors must be doubled compared to the type II string. \subsection{The NS-NS sector} The NS-NS sector is common to all of these theories and can in principle be obtained from a sigma-model approach. It involves the massless fields of the $(NS+,NS+)$ sector (a dilaton $\Phi$, a graviton $g_{MN}$ and an antisymmetric tensor $B_{MN}$) and a tachyon $T$ from the $(NS-,NS-)$ sector. The tachyon potential $V(T)$ is an even function from the property a) of the previous section~\cite{Klebanov:a}. The relevant action is thus, in the string frame \begin{equation} S_{NS-NS} = \int {\mathrm d}^dx\;\sqrt{-g}\left\{{\mathrm e}^{-2\Phi}\left( R -\frac{1}{12} |{\mathrm d} B|^2 + 4 |{\mathrm d}\Phi|^2 - \frac{1}{2}|{\mathrm d} T|^2 - V(T) \right) \right\}, \label{nspart} \end{equation} where it is natural to absorb the central charge deficit $10-d$ into the definition of the tachyon potential, i.e. \begin{equation} V(T) = -10 + d - \frac{d-2}{8} T^2 + \cdots \end{equation} It should be kept in mind that (\ref{nspart}) is by no means unique. It suffers from the usual ambiguities that come from extrapolating on-shell data. In particular, there could be arbitrary (even) functions of $T$ multiplying the various kinetic terms in the Lagrangian. Up to this order in $\alpha^\prime$ this is essentially all that can happen\footnote{Things become even more complex at the next order in $\alpha^\prime$ -- for instance there could be terms of the type $T^{2n}R^{MN}\partial_M T\partial_N T$.}. As shown in the Appendix A of~\cite{Klebanov:a} however, precisely because of their ambiguous nature it is possible to redefine away some of them, such as the term $RT^2$. At the same time, terms of the type $T^2 |{\mathrm d} B|^2$ and their counterpart for the RR kinetic terms are needed and should be kept. In the following we will never need the field $B_{MN}$ and we will set the coefficients of $R$, $|{\mathrm d}\Phi|^2$ and $|{\mathrm d} T|^2$ as in (\ref{nspart}), the main conclusions being independent of the presence of such terms. \subsection{RR kinetic and Chern--Simons terms} One guiding principle~\cite{Armoni} in the identification of the RR sector is the idea that for any $d$ there will still be massless excitations in the R sector of the open string and thus their on-shell degrees of freedom will fall into representations of the little group $SO(d-2)$. The resulting situation is best summarized in the table below. In ten dimensions one obtains massless RR fields in the type 0A theory by considering tensor products of spinors of different chiralities $(\mathbf{+-})$ and $(\mathbf{-+})$, and in the 0B theory of same chiralities $(\mathbf{++})$ and $(\mathbf{--})$. This can be readily generalized for any non-critical even dimension, whereas it is not quite clear what the right generalization to odd dimensions is. Note, however, that an odd dimensional bispinor can be decomposed in terms of the lower even dimensional bispinors as $({\mathbf ++})\oplus({\mathbf +-})\oplus({\mathbf -+})\oplus({\mathbf --})$, and that this is the sum of the field contents of the 0A and the 0B theories of one lower dimension. It thus seems reasonable to assume that a sum of two modular invariant sectors should yield a modular invariant theory in one dimension higher without doubling the RR spectrum by hand\footnote{A different point of view was taken in \cite{Armoni}. It will be interesting to find a resolution to this puzzle.}. {\footnotesize \begin{center} \begin{tabular}{c|c|c|c|c|c} $d$ & $SO(d-2)$ & Spin reps. & & $R\times R$ sector(s)& Real off-shell fields \\ &&&&&\\ \hline &&&&&\\ $4$ & $U(1)$ & ${\mathbf 1}_{1/2}$, ${\mathbf 1}_{-1/2}$ &0A& ${\mathbf 1}_0+{\mathbf1}_0$ & $2A$ \\ &&&0B& ${\mathbf 1}_{-1}+{\mathbf 1}_{1}$ & $A_M$ \\ &&&&&\\ \hline &&&&&\\ $5$ & $SU(2)$ & $\mathbf{2}$ &0AB & $\mathbf{1}+\mathbf{3}$ & $A, A_M$ \\ &&&&&\\ \hline &&&&&\\ $6$ & $SU(2)^2$ & $(\mathbf{1},\mathbf{2}),\; (\mathbf{2},\mathbf{1})$ &0A& $2(\mathbf{2},\mathbf{2})$ & $2A_M$ \\ &&&0B& $2(\mathbf{1},\mathbf{1})+(\mathbf{1},\mathbf{3})+ (\mathbf{3},\mathbf{1})$ & $2A, A_{MN}$ \\ &&&&&\\ \hline &&&&&\\ $7$ & $ Sp(4)$ & $\mathbf{4}$ &0AB& $\mathbf{1}+\mathbf{5} +\mathbf{10}$ & $A, A_M, A_{MN}$ \\ &&&&&\\ \hline &&&&&\\ $8$ & $SU(4)$ & $\mathbf{4},\; \mathbf{\bar 4}$ &0A& $2(\mathbf{1}+\mathbf{15})$ & $2A, 2A_{MN}$ \\ & & &0B& $\mathbf{6}+\mathbf{10}+\mathbf{\bar{6}}+\mathbf{\bar{10}} $ & $2A_M, A_{MNR}$\\ &&&&&\\ \hline &&&&&\\ $9$ & $SO(7)$ & $\mathbf{8}$ &0AB& $ \mathbf{1}+\mathbf{7}+\mathbf{21}+\mathbf{35}$ & $A, A_M, A_{MN}, A_{MNR}$ \\ &&&&&\\ \hline &&&&&\\ $10$& $SO(8)$ & $\mathbf{8}_s,\; \mathbf{8}_c$ &0A& $2(\mathbf{8}+\mathbf{56})$ & $2A_M, 2A_{MNR}$ \\ & & &0B& $2(\mathbf{1}+\mathbf{28}+\mathbf{35})$ & $2A, 2A_{MN}, A_{MNRP}$ \end{tabular} \end{center} } To understand the table consider for example $d=8$. In the type 0B theory, the bispinor in the $(R+,R+)$ sector decomposes into $\mathbf{4}\times\mathbf{4}=\mathbf{6}+\mathbf{10}$. These are all complex representations that yield a complex vector and a complex three form with a self dual field strength. Notice that it is possible to construct a self dual form in $d=8$ only if it is complex because $*^2=-1$. The bispinor in the $(R-,R-)$ sector yields the complex conjugate fields. These two sets of fields can be combined into two real one-forms and one real three-form without any duality constraint. These are the fields written in the last column. Notice that in odd dimensions we have only one version of the theory (0AB) without the restriction on the rank of the forms. In even dimensions the forms come in even or odd rank depending on the RR projection. In $d=10$ and $d=6$ the assignment is the familiar one (odd forms for type A and even for type B) whereas in $d=8$ and $d=4$ it is reversed. Of course, it is possible to dualize some of the fields to obtain the magnetically charged branes. The unique form of degree $d/2 -1$ for the type 0B case admits both electric and magnetic charges. Notice that if it were not for the doubling of the RR sectors it would be impossible to write off-shell real fields for the 0B theory in $d=4$ and $d=8$ because, due to the Minkowski signature, it is impossible to impose either self-duality or anti-self-duality on real forms. We view this fact, together with the argument based on modular invariance, as yet another piece of evidence for the necessity of the presence of both RR sectors. To obtain the complete form of the RR couplings up to two derivatives we need to address the issue of Chern--Simons terms. The terms of relevance here are those constructed with 3 gauge fields and 2 derivatives. Despite the lack of space-time supersymmetry, the Chern--Simons terms are present, at least in $d=10$, because they are there in the type II theories. It thus seems that the presence of these terms is dictated more by world-sheet supersymmetry than by space-time supersymmetry and it is natural to assume that such terms are also present in lower dimensions. The presence of so many RR fields may seem to lead to difficulties in determining these terms. However, there are two simplifying features that we infer from the $d=10$ case: First, there will not be terms involving only RR fields since they correspond to the correlation function of an odd number of spin fields. One of the three gauge fields must therefore be the $(NS+, NS+)$ two-form $B_{MN}$. Second, applying the selection rules of the previous section, we see that the coupling will involve fields from the same RR sector in $d=6$ and $d=10$ and from the opposite sector in $d=4$ and $d=8$. Let us start with the case of $d$ odd. In this case our proposal for the RR part of the action is \begin{eqnarray} S^{{\mathrm 0AB}}_{d=5} &=& \int\;f(T) (F_1*F_1 + F_2*F_2) + B F_1 F_2\nonumber\\ S^{{\mathrm 0AB}}_{d=7} &=& \int\;f(T) (F_1*F_1 + F_2*F_2 + \tilde F_3*\tilde F_3) + B F_2 F_3\nonumber\\ S^{{\mathrm 0AB}}_{d=9} &=& \int\;f(T) (F_1*F_1 + F_2*F_2 + \tilde F_3*\tilde F_3 + \tilde F_4*\tilde F_4) + B F_3 F_4. \label{actodd} \end{eqnarray} The forms $F_n$ are the field strengths associated to the RR gauge potentials. In this section we use ``index free'' notation and redefine the normalization coefficients to one in order not to clutter the formulas too much. A wedge product between forms is always understood. The tilde above the forms always indicates the ``NS-NS shift'', e.g. $\tilde F_4=F_4 + B F_2$, with the appropriate modified gauge transformation just as in type II supergravity. Notice that, once it is assumed that the Chern--Simons terms are present, the modification in the field strength must also be present for the action to transform correctly under electric/magnetic duality. $f(T)$ is a function of the tachyon whose first few coefficients in the Taylor expansion around zero can in principle be determined by extrapolating from the on-shell computation~\cite{Klebanov:a} in $d=10$. We shall see that the detailed form of these functions is not directly relevant for the computations of the properties of the dual field theories. To write the actions for the even dimensional cases, let us denote the field strengths from the two RR sectors by $F, \; F^\prime$ for $d=6$ and $d=10$ and by $F, \; \bar F$ for $d=4$ and $d=8$. In the former case the field strengths are real whereas in the latter they are complex conjugates of each other. The form of highest degree in the type 0B case is special -- it is self dual in the complex case and its vertex operator does not contain the chiral projection. { \begin{eqnarray} S^{{\mathrm 0A}}_{d=4} &=& \int\;f_{\mathrm even}(T) (F_1*\bar F_1) + f_{\mathrm odd}(T) (F_1*F_1+ \bar F_1*\bar F_1) +i B F_1 \bar F_1\nonumber \\ S^{{\mathrm 0B}}_{d=4} &=& \int\;f_{\mathrm even}(T) (F_2*\bar F_2) + f_{\mathrm odd}(T) (F_2*F_2 +\bar F_2*\bar F_2) \nonumber \\ S^{{\mathrm 0A}}_{d=6} &=& \int\;f_{\mathrm even}(T) (F_2* F_2 + F^\prime_2* F^\prime_2) + f_{\mathrm odd}(T) (F_2* F^\prime_2) + B (F_2 F_2 + F^\prime_2 F^\prime_2) \nonumber \\ S^{{\mathrm 0B}}_{d=6} &=& \int\;f_{\mathrm even}(T) (F_1* F_1 + F^\prime_1* F^\prime_1+ \tilde F_3 *\tilde F_3 ) + f_{\mathrm odd}(T) (F_1* F^\prime_1 + \tilde F_3 *\tilde F_3 ) \nonumber\\ && + B (F_1 F_3 + F^\prime_1 F_3) \nonumber \\ S^{{\mathrm 0A}}_{d=8} &=& \int\;f_{\mathrm even}(T) (F_1*\bar F_1 + \tilde F_3 * \tilde {\bar F_3}) +f_{\mathrm odd}(T) (F_1*F_1+ \bar F_1*\bar F_1 + \tilde F_3 * \tilde F_3 \nonumber \\ &&+\tilde{\bar F_3} * \tilde{\bar F_3} ) +i B F_3 \bar F_3\nonumber \\ S^{{\mathrm 0B}}_{d=8} &=& \int\;f_{\mathrm even}(T) (F_2*\bar F_2 + \tilde F_4 * \tilde{\bar F_4}) +f_{\mathrm odd}(T) (F_2*F_2 +\bar F_2*\bar F_2 \nonumber\\ &&+ \tilde F_4 * \tilde F_4 +\tilde{\bar F_4} * \tilde{\bar F_4}) + B (F_2 \bar F_4 + F_4 \bar F_2) \nonumber \\ S^{{\mathrm 0A}}_{d=10} &=& \int\;f_{\mathrm even}(T) (F_2* F_2 + F^\prime_2* F^\prime_2 + \tilde F_4* \tilde F_4 +\tilde F^\prime_4*\tilde F^\prime_4) \nonumber \\&&+ f_{\mathrm odd}(T)( F_2* F^\prime_2 + \tilde F_4*\tilde F^\prime_4) + B (F_4 F_4 + F^\prime_4 F^\prime_4) \nonumber \\ S^{{\mathrm 0B}}_{d=10} &=& \int\;f_{\mathrm even}(T) (F_1* F_1 + F^\prime_1* F^\prime_1+ \tilde F_3* \tilde F_3 + \tilde F^\prime_3* \tilde F^\prime_3 + \tilde F_5 *\tilde F_5 )\nonumber \\&& + f_{\mathrm odd}(T) (F_1* F^\prime_1 + \tilde F_3 *\tilde F^\prime_3 + \tilde F_5 *\tilde F_5 ) + B (F_3 F_5 + F^\prime_3 F_5).\label{acteven} \end{eqnarray} } We present the form of the actions in all their generality because it will be useful for future more detailed computations. For our present purposes however, it should be noticed that the kinetic terms in the actions can be diagonalized by letting $F_\pm = F\pm F^\prime$ in $d=6,10$ and $F_\pm = F\pm i\bar F$ in $d=4,8$. \subsection{Massive type 0 gravity} There is still one RR form field that can be added to the actions (\ref{actodd}) and (\ref{acteven}). In a $d$-dimensional space-time it is possible to introduce a rank $d-1$ gauge potential coupling to a corresponding extended object. It carries no physical degrees of freedom and therefore it is not visible in the on-shell analysis of the previous subsection. Its rank $d$ field strength, however, carries an energy density and it does affect the physics. This form will be used in Section 5 when constructing various field theory duals. This case will provide the simplest example which displays most of the interesting physics, and it allows one to avoid the complications of disentangling the Kaluza-Klein modes. In type IIA supergravity (and thus in $d=10$ type 0A for each RR sector) it is well known how to introduce such a field~\cite{Romans,Bergshoeff:1996ui}. The required modifications in the bosonic sector are the addition of the terms \begin{eqnarray} \int\; M F_{10} +\frac{1}{2} M^2 *1 \label{addit} \end{eqnarray} to the action, and a further shift of the 2- and the 4-form field strengths by $MB$ and $MB^2/2$, respectively. The gauge transformations are changed accordingly in order to re-ensure gauge invariance. Integrating over the gauge potential of $F_{10}$ imposes the constraint that $M$ be constant. Solving the equation of motion for $M$ establishes a connection between $M$ and $F_{10}$. In the case $B=0$ -- relevant to our analysis -- they are simply Hodge duals of each other as is readily seen from (\ref{addit}). From the string theory point of view \cite{Polchinski:b}, the natural generalization of the RR $\beta$-function equations implies ${\mathrm d} * F_{d}=0 $ and ${\mathrm d} F_{d}=0$, as the top-form, too, appears in the reduction of an even dimensional type 0A bispinor into antisymmetric tensor representations. In our case, we must also include the coupling with the tachyon. The relevant addition is \begin{eqnarray} - \frac{1}{2} \int f(T) F_{d} * F_{d} \end{eqnarray} that we assume be present in any dimension. \section{Classical solutions} In what follows we shall show that the above described low energy theories allow Freund--Rubin type solutions~\cite{Freund}, where the dilaton and the tachyon are constant, the space-time factorizes into a product of an AdS space and a sphere, and the only nontrivial form-field is a RR field. Such types of solutions are very familiar from the supergravity literature, see e.g.~\cite{Duff:a, Duff:b, Kim, Salam, Randjbar-Daemi}. It is sufficient to consider the Einstein frame action\footnote{We now switch to component notation for clarity and reinstate all the appropriate normalizations. Note that our normalization of the tachyon differs by a factor of $\sqrt{2}$ from that of most of the recent literature.} \begin{eqnarray} S &=& \int {\mathrm d}^dx \sqrt{-g}~ \Bigg\{ R - \frac{1}{2} (\partial_{M}\Phi)^2 - \frac{1}{2} (\partial_{M}T)^2 - V(T)~{\mathrm e}^{a\Phi} \nonumber \\ & & \qquad - \frac{1}{2~(p+2)!}~ f(T)~ {\mathrm e}^{b\Phi}~ \Big(F_{M_1 \cdots M_{p+2}}\Big)^2 \Bigg\}, \end{eqnarray} where $V(T)$ is the sum of the tachyon potential and the central charge deficit, and $f(T)$ is the coupling between the $(p+2)$-dimensional RR form $F$ and the tachyon. The RR gauge field is the appropriate linear combination of some of the fields of the previous section in such a way that the kinetic terms are diagonal. After diagonalization, $f(T)$ no longer has any particular symmetry property. The coefficients $a$ and $b$ are \begin{eqnarray} a &=& \sqrt{\frac{2}{d-2}} \\ b &=& \frac{1}{2} (d-2p-4) \sqrt{\frac{2}{d-2}} . \end{eqnarray} The field $B_{MN}$ that we are setting to zero here may appear linearly in the full action only in the Chern--Simons term, but in that case multiplied by $F \wedge F$, which will vanish in the Freund--Rubin ansatz. The equations of motion can be summarized as follows: \begin{eqnarray} \Box \Phi &=& aV(T)~ {\mathrm e}^{a\Phi} + \frac{b}{2}~ f(T)~ {\mathrm e}^{b\Phi}~ \frac{1}{(p+2)!}~ \Big(F_{M_1 \cdots M_{p+2}}\Big)^2 \label{dilaton} \\ \Box T &=& V'(T)~ {\mathrm e}^{a\Phi} + \frac{1}{2}~ f'(T)~ {\mathrm e}^{b\Phi}~ \frac{1}{(p+2)!}~ \Big(F_{M_1 \cdots M_{p+2}}\Big)^2 \label{tachyon} \\ R_{MN} &=& \frac{1}{2}~ \partial_M \Phi~ \partial_N \Phi + \frac{1}{2}~ \partial_M T~ \partial_N T \nonumber \\ & & + \frac{1}{d-2}~ g_{MN}~ V(T)~ {\mathrm e}^{a\Phi} + \frac{1}{2}~ f(T)~ {\mathrm e}^{b\Phi}~ \tilde{T}_{MN} \label{gravity} \\ 0 &=& \nabla^N~ \Big( f(T)~ {\mathrm e}^{b\Phi}~ F_{NM_1 \cdots M_{p+1}} \Big)~. \label{formfield} \end{eqnarray} The tensor $\tilde T_{MN}$ is shorthand for the (trace subtracted) stress energy tensor \begin{eqnarray} \tilde{T}_{MN} &=& \frac{1}{(p+1)!} \Bigg( F_{MK_1 \cdots K_{p+1}} F_{N}^{~~K_1 \cdots K_{p+1}} - \frac{(p+1)~ g_{MN}}{(p+2)(d-2)} \Big(F_{K_1 \cdots K_{p+2}}\Big)^2 \Bigg) \nonumber \\ & & \end{eqnarray} Again, we have ignored potential contributions from Chern--Simons terms because they vanish for the classical solution. They do contribute to the analysis of the fluctuations in the general case and also for this reason, in the next section, when computing the critical properties of the field theory duals we restrict to the simple case $d=p+2$ where such complications do not arise. We hope to return to the general case in a later paper~\cite{Ferretti:b}. These equations of motion have a solution with constant dilaton $\Phi=\Phi_0$ and tachyon $T=T_0$ in the gravity background of a product space \begin{eqnarray} {\mathrm AdS}_{p+2} \times {\mathrm S}^{d-p-2}. \end{eqnarray} The size of the two maximally symmetric spaces is determined by setting\footnote{The Greek indices refer to the AdS space and the Latin indices to the sphere.} (always in units of $\alpha^\prime$) \begin{equation} R_{\mu\nu\rho\lambda}=-\frac{1}{R^2_0}\left(g_{\mu\rho}g_{\nu\lambda} -g_{\mu\lambda}g_{\nu\rho} \right)\quad R_{ijkl}=+\frac{1}{L^2_0}\left(g_{ik}g_{jl}-g_{il}g_{jk} \right). \label{riemann} \end{equation} Finally, the RR field is set proportional to the volume-form of the anti-de Sitter space, and hence its only nontrivial components are \begin{eqnarray} F_{\mu_1 \cdots \mu_{p+2}} &=& F_0~ \sqrt{-g({\mathrm AdS}_{p+2})}~ \epsilon_{\mu_1 \cdots \mu_{p+2}}, \end{eqnarray} where the constant $F_0$ is related to the conserved charge $k$ by \begin{eqnarray} k = f(T_0)~{\mathrm e}^{b\Phi_0}~F_0~. \label{chargek} \end{eqnarray} Given the two functions $V(T)$ and $f(T)$, the tachyon and the dilaton vacuum expectation values are determined from Eqs.~(\ref{dilaton}), (\ref{tachyon}) and (\ref{chargek}). The tachyon $T_0$ can be expressed implicitly, as the solution of an algebraic equation, namely \begin{eqnarray} \frac{f'(T_0)}{f(T_0)} &=& \frac{1}{2}(d-2p-4)~ \frac{V'(T_0)}{V(T_0)}. \label{tnaught} \end{eqnarray} The dilaton $\Phi_0$ can then be readily obtained from \begin{eqnarray} {\mathrm e}^{(a+b)\Phi_0} &=& \frac{(d-2p-4)}{4}~ \frac{k^2}{f(T_0)~V(T_0)}~. \end{eqnarray} The radii of the anti-de Sitter space $R_0$ and that of the sphere $L_0$ can be solved from the Einstein equations (\ref{gravity}) \begin{eqnarray} R^2_0 &=& (p+1)(d-2p-4) \frac{{\mathrm e}^{-a\Phi_0}}{V(T_0)} \label{rrr} \\ L^2_0 &=& (d-p-3)(d-2p-4) \frac{{\mathrm e}^{-a\Phi_0}}{V(T_0)}. \label{landlambda} \end{eqnarray} In the derivation we assumed $k\neq 0$. Also three special dimensionalities were excluded for compactness: \begin{itemize} \item[a)] The case $d = p+3$ leads to an infinite radius in the AdS space-time, i.e.~the flat Minkowski space times a circle, and is not considered in what follows. \item[b)] For $d=2p+2$ the dilaton becomes a free parameter. Rather than its vacuum expectation value $\Phi_0$, the charge $k$ is determined from the equation of motion (\ref{dilaton}) \begin{eqnarray} k^2 = -2 V(T_0)~f(T_0)~ . \end{eqnarray} The rest of the formulae (\ref{tnaught}), (\ref{rrr}) and (\ref{landlambda}) are still valid. \item[c)] We assumed that $V(T_0) \neq 0$. In addition to some completely Ricci flat solutions this condition also excludes the middle dimensional branes, for which we have $d=2p+4$. In these dimensions the radii are \begin{eqnarray} R_0^2 = L_0^2 = 4~(p+1)~ \frac{f(T_0)}{k^2}~, \end{eqnarray} Now $T_0$ is determined from $V(T_0) = 0$ (not $V^\prime(T_0)=0$) and $\Phi_0$ from \begin{eqnarray} V'(T_0) {\mathrm e}^{a\Phi_0} - \frac{k^2}{2} \frac{f'(T_0)}{f(T_0)^2} =0~. \end{eqnarray} \end{itemize} The solutions discussed above are physically acceptable only for $f(T_0)>0$. \section{Dual field theory interpretation} In this section we finally make contact with the conjectured gravity/field theory duality by studying the field theory duals of the type 0 theories for the simple case of $d=p+2$. This case already contains all the relevant qualitative features of the most general one, without the complication of the Kaluza--Klein analysis. Let us state the logic of the approach. The classical solutions on the gravity side correspond to fixed points on the field theory side. There is a geodesic flow that relates these classical solutions, which is interpreted as the renormalization group flow connecting different fixed points. Fluctuation modes with positive, vanishing, and negative mass square correspond to irrelevant, marginal, and relevant deformations. The critical exponents can be obtained from these masses and they depend on a finite set of undetermined parameters due to the arbitrariness of the tachyon couplings. These parameters should be fixed by comparing some universal quantities with experiment which leads to a prediction for the remaining quantities. \subsection{Stability of the AdS$_d$ solutions} Classical solutions can serve as sound vacua for a quantum theory only if small fluctuations around the solutions are stable. In Minkowski space this implies that tachyonic fluctuation modes are forbidden. In an AdS background this requirement can be relaxed, and one finds the bound \cite{Abbott, Breitenlohner, Mezincescu:a, Mezincescu:b} \begin{eqnarray} m^2 \geq - \frac{(d-1)^2}{4}~ \frac{1}{R_0^2}~. \label{BF} \end{eqnarray} for the masses of the scalar fluctuation modes. The first source of these instabilities near the solutions found in the previous section are obviously fluctuations in the tachyon field $T$. Tachyonic instabilities may enter also through the various scalar fields that appear on the AdS space, as the fields are compactified on the sphere $S^{d-p-2}$. In order to show that the theory is stable against these perturbations, one has to linearize the full set of equations of motion around the classical solution, and check that no mode violates the bound (\ref{BF}). This can be done, but the physically relevant features already appear in the case where the transverse sphere is absent. We shall discuss this example below in detail. As $d=p+2$, the nontrivial RR field is the top-form, dual to a cosmological constant \begin{eqnarray} F_{\mu_1 \cdots \mu_{p+2}} &=& F~ \sqrt{-g}~ \epsilon_{\mu_1 \cdots \mu_{p+2}}~. \end{eqnarray} Note, that this is no longer the Freund--Rubin ansatz, but the RR field is a priori entirely general and unconstrained. This is the field discussed in Section 3.3. The equation of motion (\ref{formfield}) becomes in this case a constraint, and it turns out that the conserved charge is \begin{eqnarray} k = f(T)~{\mathrm e}^{b\Phi}~F~. \label{cha} \end{eqnarray} With the help of (\ref{cha}), the equations of motion for the other fields reduce to a Hamiltonian form \begin{eqnarray} \Box \Phi &=& -\frac{\partial}{\partial\Phi} {\cal V}(\Phi,T) \\ \Box T &=& -\frac{\partial}{\partial T} {\cal V}(\Phi,T) \\ R_{\mu\nu} &=& \frac{1}{2}\partial_\mu\Phi\partial_\nu\Phi+\frac{1}{2}\partial_\mu T\partial_\nu T-\frac{1}{d-2}{\cal V}(\Phi,T)g_{\mu\nu} \end{eqnarray} where the effective potential is \begin{equation} {\cal V}(\Phi,T) = - V(T) {\mathrm e}^{a\Phi} - \frac{1}{2}~ \frac{k^2}{f(T)}~ {\mathrm e}^{-b\Phi}~. \end{equation} Let us linearize the equations of motion near a classical solution \begin{eqnarray} \Phi &=& \Phi_0 + \varphi \\ T &=& T_0 + t \\ g_{\mu\nu} &=& \hat{g}_{\mu\nu} + h_{\mu\nu}~. \end{eqnarray} In order to do this, we need some knowledge of the functions $V(T)$ and $f(T)$. The only characteristics of these functions that will enter the stability analysis are the coefficients \begin{eqnarray} x = \frac{V'(T_0)}{V(T_0)}~, \qquad y = \frac{V''(T_0)}{V(T_0)}~, \qquad \mbox{and} \qquad z = \frac{f''(T_0)}{f(T_0)} \label{xxyyzz}~. \end{eqnarray} Perturbative string theory analysis around $T=0$ yields \cite{Klebanov:a} \begin{eqnarray} V(T) &=& d-10 - \frac{d-2}{8}~T^2 + {\cal O}(T^4) \\ f(T) &=& 1 + T + \frac{1}{2}~ T^2 + {\cal O}(T^3). \end{eqnarray} This is not enough to determine the coefficients (\ref{xxyyzz}), and they should indeed be treated as free parameters of the theory. Including other unknown functions (see discussion in Section 3.1) would give rise to more than three such parameters, but the analysis performed here would still have the same qualitative features. The fact that the graviton fluctuations actually decouple completely from those of the scalars simplifies the calculations: The graviton equations of motion can, in fact, be derived to first order from the effective action \begin{eqnarray} S_{{\mathrm h}} &=& \int {\mathrm d}^dx~ \sqrt{-(\hat{g}+h)}~ \Bigg\{ R(\hat{g}_{\mu\nu} + h_{\mu\nu}) + {\cal V}(\Phi_0,T_0) \Bigg\}~. \end{eqnarray} The scalar fluctuations obey \begin{eqnarray} \Big( -{\Box}+ {\cal M}~ \Big) \left(\begin{tabular}{c} $\varphi$\\ $t$ \end{tabular} \right) =0 \label{scalfluct} \end{eqnarray} where the mass matrix is \begin{eqnarray} {\cal M} = d(d-1)~ R^{-2}_0~ \left(\begin{tabular}{cc} $1$ & $\sqrt{\frac{d-2}{2}}\,x$ \\ $\sqrt{\frac{d-2}{2}}\,x$ & $d~ x^2 -y - \frac{2}{d} z$ \end{tabular} \right)~. \end{eqnarray} The mass eigenvalues are \begin{eqnarray} m^2_{1,2} &=& d(d-1) \, R^{-2}_0 \, \Bigg(1 + \frac{\tau}{2} \pm \frac{1}{2}\sqrt{\tau^2+(2d-4) x^2}\Bigg) \end{eqnarray} where \begin{eqnarray} \tau = d~ x^2 -\frac{2z}{d}-y-1~. \end{eqnarray} Note that the masses depend only on two independent parameters $x$ and $\tau$. If we assume, following \cite{Klebanov:b}, that $f(T) = \exp(T)$, then the equations of motion give $x=-2/d$, and we can easily extract some interesting qualitative features as the only undetermined parameter is $\tau$. In this case there turns out to be three different, continuously connected phases: First, there can be two particles, both with positive mass squared. Second, there can be a particle and a tachyon that obeys the bound (\ref{BF}). Third, there can be a tachyon that makes the vacuum unstable. In AdS/CFT correspondence this translates into the statement that there can be at most one relevant operator in the infrared near the fixed point described by this theory. \subsection{Solutions connecting conformal fixed points} The stability analysis as applied to the critical points of the potential yields local information about the behavior of the dual field theory near its fixed points. Depending on the form of the potential, there may exist gravity solutions that interpolate between different critical points. These solutions should be interpreted on the field theory side as RG trajectories between conformal points. In order to study these interpolating solutions we consider the ansatz \begin{equation} \label{Liouv-ansatz} {\mathrm d} s^2={\mathrm d} y^2+A^2(y)~ {\mathrm d} x^2_{\parallel} \end{equation} and allow the two scalars to depend on the Liouville coordinate $y$. We already know from the previous sections that there are exact solutions of the form \begin{eqnarray} A(y) = {\mathrm e}^{y/R}~, \end{eqnarray} where $R$ is the radius of the pertinent AdS space. The Einstein equation gives rise to two independent equations. Defining the following auxiliary function \begin{equation} \gamma(y)=(d-1)~ \frac{{\mathrm d}}{{\mathrm d} y}\log (A)~, \end{equation} the full set of equations takes the form \begin{eqnarray} \frac{\ddot{A}}{A} + (d-2)\left( \frac{\dot{A}}{A} \right)^2 &=& \frac{1}{d-2}{\cal V} \label{dump1}\\ \ddot{\vec{{\Phi}}}+\gamma\dot{{\vec{\Phi}}} &=& -\vec{\nabla}{\cal V} \label{dump3} \\ \dot{\gamma} &=& -\frac{d-1}{2(d-2)}\big(\dot{\vec{\Phi}}\big)^2 \leq 0 \label{dump2}~. \end{eqnarray} Here we denote derivatives with respect to $y$ with a dot, and we have introduced the compact notation $\vec{\Phi}=(\Phi,T)$ for the two scalars. Provided $\gamma \geq 0$, equation (\ref{dump3}) has the physical interpretation of a particle moving on a plane in the potential ${\cal V}$, subject to a friction force. Let us assume that the potential has two critical points $\vec{\Phi}_1$ and $\vec{\Phi}_2$ that satisfy ${\cal V}(\vec{\Phi}_1) > {\cal V} (\vec{\Phi}_2)$, and that there is at least one unstable direction at $\vec{\Phi}_1$ for increasing $y$ and, similarly, a stable direction for $\vec{\Phi}_2$. This can always be arranged by choosing the ${\cal O}(T^4)$ part in $V(T)$ suitably. Due to the friction coefficient we expect our particle to roll down starting from the IR fixed point, and to converge in an infinite amount of time towards the lower UV fixed point. This happens, since $\gamma$ is strictly positive: Indeed, at the critical points $\gamma$ approaches the values \begin{eqnarray} \gamma &\to& \frac{d-1}{R_1} \quad \mbox{for } y\to -\infty \\ \gamma &\to& \frac{d-1}{R_2} \quad \mbox{for } y\to +\infty~, \end{eqnarray} and the friction coefficient decreases monotonously between them according to (\ref{dump2}). This is consistent with the fact that \begin{equation} R^2_{1,2}=\frac{(d-1)(d-2)}{{\cal V}(\vec{\Phi}_{1,2})} \end{equation} as follows from (\ref{dump1}). The solution might be oscillatory near the UV critical point. Whether this happens depends on whether the friction is enough to stop the particle as it arrives at the lower point. Clearly, if one wants to interpret the result as an RG flow, the oscillatory behavior would be difficult to accommodate in the field theory picture. The oscillatory solutions are exactly the solutions that would violate the bound (\ref{BF}), which is a necessary condition for the consistency of the system on the gravity side. Hence, quite remarkably, the stability in the gravity theory is dual to the consistency of the field theory interpretation. Some universal information can be read from the local behavior of these solutions. In the spirit of the Wilsonian RG treatment, let us study the critical behavior near the two fixed points in the linearized approximation. We must first identify the appropriate coordinate which in field theory can be consistently interpreted as the energy scale and parameterizes the interpolating solution. Such a coordinate can be chosen to be\footnote{This definition corresponds locally, near the fixed points, to the one used in \cite{Maldacena}. However, there are alternative definitions. For instance choosing $U=\dot A$ one obtains the holographic relation, cf.~\cite{Peet}. All of these definitions lead to the same universal quantities.} \begin{equation} U=\frac{A^2}{\dot{A}} \end{equation} since at the critical points this reduces to $U=R~ {\mathrm e}^{y/R}$, where the metric takes the standard form \begin{equation} {\mathrm d} s^2 = \frac{R^2}{U^2} {\mathrm d} U^2+\frac{U^2}{R^2} {\mathrm d} x_{\parallel}^2~. \end{equation} Define \begin{equation} \vec{\Phi}(U) = \vec{\Phi}_0+\delta\vec{\Phi}(U)~, \end{equation} so that Eq.~(\ref{scalfluct}) takes the form \begin{equation} \left[-\frac{1}{R^2}[d~U\partial_U+U^2\partial_U^2]+ {\cal M}\right]\delta\vec{\Phi}=0. \end{equation} The eigenvalues of $\cal M$, namely $m_i^2$ are found in Section 5.1, and for each of the two eigenvectors we get two linearly independent solutions \begin{equation} \delta\tilde{\Phi}_i=A_iU^{\lambda^i_+}+B_iU^{\lambda^i_-}~, \end{equation} where \begin{equation} \lambda^i_\pm=\frac{-(d-1)\pm\sqrt{(d-1)^2+4m_i^2R^2}}{2}~. \label{roots} \end{equation} Notice first that the stability condition (\ref{BF}) ensures the reality of the roots and they only depend on the dimensionless parameters $x$, $y$, and $z$. The IR limit corresponds to taking $U\to 0$, for which there must be at least one positive eigenvalue, say, $(m^{{\mathrm IR}}_1)^2 >0$. In order for the solution not to blow up at this point we must choose $B^{\mathrm IR}_1=B^{\mathrm IR}_2=0$. If $(m^{{\mathrm IR}}_2)^2 <0$ we must also set $A^{\mathrm IR}_2=0$, otherwise, the trajectory may in general start with a linear combination of the two eigenvectors. The trajectory will then evolve to the UV fixed point as $U\to \infty$ where there will be at least one negative mass eigenvalue, say $(m^{{\mathrm UV}}_1)^2$. Generically, both the coefficients $A^{\mathrm UV}_1$ and $B^{\mathrm UV}_1$ will be non-zero and the root $\lambda^1_+$ will dominate. In the AdS/CFT correspondence it is common to identify $g_{\mathrm YM}^2={\mathrm e}^\Phi$. This relation is plausible by considering the weak coupling expansion of the world-volume gauge theory living on the stack of D-branes. On the other hand, away from the Gaussian fixed point there is no a priori reason to make such an identification. Thus we will stay general and regard $\Phi$ and $T$ as the coupling constants. We can now read off the leading order behavior of the $\beta$-functions near the fixed points: \begin{equation} \beta_i(g_i)=U\frac{{\mathrm d} g_i}{{\mathrm d} U}=\lambda_+^i(g_i-g_i^*)+\dots, \end{equation} where $g_i=\tilde{\Phi}_i$. In particular, the conformal dimension of an operator coupled to the bulk field $\tilde\Phi_i(U,x)$ (a linear combination of the original tachyon and dilaton) is $\Delta_i= d-1 + \lambda_+^i$ and the anomalous dimension is $\lambda_+^i$. Note that, if $\lambda_+$ vanishes, then we should have $A=0$ as well. This represents a marginal operator within our approximation and the computation of the $\beta$-function would pick up the sub-leading contribution which in the UV is $\lambda_- = -(d-1)$. This is what happens in~\cite{Kehagias:a}, where the RG flow is studied in the framework of type II supergravity. One could also study confining and asymptotically free solutions of these models as well as extend to situations where Kaluza--Klein modes are present. We hope to return to some of these issues in the future. \section{Acknowledgments} We wish to thank R. Iengo, G. Mussardo, S. Randjbar-Daemi and A. Schwimmer for discussions. We are also grateful to all the participants to the A. Salam I.C.T.P. journal club for providing a forum for discussion. This work was supported in part by the European Union TMR programs CT960045 and CT960090.
\section{Introduction} The recent run (1997-98) of the Stanford Linear Collider (SLC) was the most productive to date : approximately 350K Z bosons were detected, compared to 210K for the entire program from 1992-96, at peak luminosities of $3 \times 10^{30} {\rm cm^{-2} s^{-1}}$, nearly a factor of three improvement compared to the best previous results. Nevertheless, LEP enjoys a 28 : 1 advantage in statistics, and it is only due to the unique features of SLC operation that the SLD experiment is able to contribute several state-of-the-art electroweak and b-physics measurements. These well-known features are : \begin{itemize} \item High (75\%), precisely measured (${{\delta \cal P} \over{P}} \sim 0.5\%$) longitudinal e- polarization. \item A small and stable $e^+e^-$ luminous region (1.5 by 0.8 by 700 $\mu m$) and a uniquely precise CCD-based vertex detector (I.P. determined to 4 by 4 by 30 $\mu m$). \end{itemize} In what follows, recent electroweak results will be summarized, some historical background provided, and implications of the data will be discussed. \section{The Electroweak Observables} The polarized differential cross section at the Z pole is given by : $$ {d\sigma \over {d cos\theta}} \sim (1 - {\cal P}_e A_e)(1 + cos^2\theta) + 2A_f(A_e - {\cal P}_e)cos\theta, $$ where the parity violating asymmetries in terms of the vector and axial vector NC couplings for fermion flavor $f$ are $A_f = {2v_f a_f \over {v_f^2 + a_f^2}}$. The polarized $e^-$ beam at the SLC allows for the isolation of the initial state ($A_e$) and final state ($A_f$) asymmetries. The initial state couplings are determined most precisely via the left-right Z production asymmetry $$ A_{LR}^0 = {1 \over {\cal P}_e} {\sigma_L - \sigma_R \over {\sigma_L + \sigma_R}} = A_e , $$ while the left-right-forward-backward asymmetry for the final state flavor $f = b,c,s,e,\mu ,\tau$ $$ A_{LRFB} = {(\sigma_{LF} - \sigma_{LB}) - (\sigma_{RF} - \sigma_{RB}) \over {(\sigma_{LF} + \sigma_{LB}) + (\sigma_{RF} + \sigma_{RB})}} = {3 \over{4}} {\cal P}_e A_f , $$ determines the final state couplings. The $A_{LR}$ measurement is unique among all electroweak precision measurements in that no efficiency or acceptance effects enter, and no significant final state identification is required. Due to the extensively crosschecked high precision polarimetry, the total systematic error ($\sim 0.75\%$) ensures that the result is statistically dominated (stat. error $\sim 1.3\%$). The quantity $A_{LR}^0$ provides by far the most precise determination of $sin^2\theta_W^{eff}$ presently available, and rivals the 4 experiment CERN average, without recourse to the assumption of lepton/hadron universality inherent in the most precise technique from LEP ($A_{FB}(b)$). The significance of the $A_{LRFB}$ measurement, while not as precise as $A_{LR}$, is that it provides the only direct measurement of the important parameter $A_b$ (and the charm, strange and muon analogs as well), which can only be obtained indirectly from unpolarized asymmetries. While the weak mixing angle measurements are particularly sensitive to vacuum polarization loop effects (and hence to the Higgs mass), $A_b$ is instead affected by corrections at the Zb${\rm \overline{b}}$ vertex. In the context of the Minimal Standard Model (MSM), these vertex corrections are insensitive to $M_{Higgs}$ and hence $A_b$ has an unambiguous predicted value (compared to experimental precision). The combination of independent measurements of $A_e$ and $A_b$ is therefore a powerful test of the MSM. In addition, measurements of the hadronic partial width ratios $R_b$ and $R_c$, which are best measured at LEP and SLD, respectively, have become precisely known. In particular, $R_b$ is interesting due to high precision (0.4\% in the world average), and the fact that it provides a nicely complimentary measurement to $A_b$ : $A_b$ is primarily sensitive to {\em right}-handed NC b couplings, while $R_b$ is most sensitive to the {\em left}-handed sector. \section{Remarks on High Precision } The unique precision of $A_{LR}$ is a centerpiece of the SLD program, but the extensive crosschecks which have bolstered our confidence in this measurement are not so well known and are briefly reviewed here. In the early years (1992-95), a number of dedicated accelerator experiments were performed to establish the integrity of the polarimetry, in particular 1) the $e^-$ bunch helicity transmission was verified by setting up a current/helicity correlation in the SLC, 2) medium precision M\o ller and Mott polarimeters confirmed the high precision Compton polarimeter result to $\sim 3\%$. In addition, the advent of spin manipulation via ``spin bumps" in the SLC arcs allowed us to minimize the spin chromaticity ($d{\cal P}/dE$) which helped reduce a resulting polarization correction from $>1\%$ in 1993 to $<0.2\%$ by 1995. Since 1997, two additional detectors of the Compton scattered photons (the Compton $e^-$ are seen in the primary device), with rather different systematics, presently confirm our overall polarization scale to within $0.5\%$. Most recently, two longstanding questions were answered : 1) A dedicated experiment using the End Station A fixed target polarimeter confirmed that accidental $e^+$ polarization is consistent with zero ($-0.02\pm0.07\%$), 2) A short resonance scan was used to calibrate the SLC energy spectrometers against $M_Z$, verifying their accuracy on $E_{cm}$ to about 40 MeV and leading to an estimate of induced systematic error of $\sim 0.5\%$. \footnote{This result was somewhat inflated by instrumental problems during the scan, compared to our prior estimate of $\sim 0.4\%$, but remains at or below the polarimeter uncertainty.} In summary, several years of instrumental work and crosschecks, supplemented by extensive accelerator based tests, have answered a large number of detailed questions, from the most fundamental to the fairly obscure. The high precision of $A_{LR}$ is now very well established. \section{Results and Interpretation} The preliminary results are given below (with the exception of kaon tagging for $A_b$, and the latest $\sim 100K$ events for $R_b$, these results are based on the entire 1992-1998 SLD data set). \cite{mor99} \begin{table}[ht] \caption{SLD electroweak results.\label{tab:exp}} \vspace{0.2cm} \begin{center} \footnotesize \begin{tabular}{|c|c|c|l|} \hline { Observable } & {Prelim. Result} & {$sin^2\theta_W$} & {comments}\\ \hline \hline { $A_{LR}$ } & {$0.1504\pm0.0023$} & {$0.23109\pm0.00029$} & {Incl. SLD leptonic result}\\ \hline \hline { $A_{e}$ } & {$0.1504\pm0.0072$} & {} & {(The LEP leptons only}\\ { $A_{\mu}$ } & {$0.120\pm0.019$} & {($0.2317\pm0.0008$)} & {result for $sin^2\theta_W$ :}\\ { $A_{\tau}$ } & {$0.142\pm0.019$} & {} & {$0.23153\pm0.00034$)}\\ \hline { $A_c$ } & {$0.634\pm0.027$} & {} & {(LEP : $0.634\pm0.040$)}\\ { $A_b$ } & {$0.898\pm0.029$} & {} & {(LEP : $0.887\pm0.021$)}\\ \hline { $R_c$ } & {$0.169\pm0.006$} & {} & {These observables are}\\ { $R_b$ } & {$0.2159\pm0.0020$} & {} & {consistent with the MSM.}\\ \hline \end{tabular} \end{center} \end{table} A few comments are in order : \begin{itemize} \item Final errors will be $\sim \pm 0.00025$ for $sin^2\theta_W^{eff}$, and $\sim \pm 0.022$ for $A_b$, mainly due to improved systematics. \item Additional $sin^2\theta_W^{eff}$ information derives from the left-right asymmetry in the lepton sample (the dominant $A_{LR}$ result is from a hadronic sample). \item The $A_b$ result is just over one sigma away from the SM prediction (0.935), but if $A_b$ is deduced from the LEP $A_{FB}(b)$ measurement, the SLD/LEP combined result is 2.6 sigma low. \item The SLD $sin^2\theta_W^{eff}$ result is nicely consistent with lepton-based results from LEP (0.8 sigma), a situation that has held stably since 1995, while the $A_{FB}(b)$ dominated LEP hadronic average differs from the lepton based result by 2.2 sigma. \end{itemize} The difficulty seen with the b-flavor results may be a statistical fluctuation or due to analysis bias, or may point to the intruiging possibility of an anamoly in b-NC couplings, in particular, the right-handed coupling, a situation which is difficult to motivate theoretically (the $R_b$ world average is consistent with the MSM). In either of the later two cases, the b-hadron based $sin^2\theta_W^{eff}$ result is called into question - for now we feel it is reasonable to perform our electroweak fits using the lepton-based weak mixing angle average. We first work within the framework of the MSM - Figure 1 shows the result of separate fits to the Higgs mass using the lepton-based $sin^2\theta_W^{eff}$ results from SLD and LEP, the $M_W$ results from LEP II and the Tevatron, and for comparison the result using the $A_{FB}(b)$ based weak mixing angle measurement from LEP. \footnote{The $\alpha(M_Z^2)$ of Kuhn etal. is used for the fits discussed here - the Jegerlehner etal. value used by the LEP EWWG yields a $\chi^2$ minimum about 30 GeV lower, but due to larger errors, provides about the same 95\% confidence upper bound.\cite{alpha}} The $M_W$ measurements seem to be confirming the very low Higgs mass favored by $A_{LR}$. It is also noteworthy that even when $M_W$ precision reaches $\pm 30$ MeV (presently $\pm 44$ MeV), the strongest contraints will still be coming from $sin^2\theta_W^{eff}$. The key to this enterprise is that improved $\alpha(M_Z^2)$ determinations are becoming available \cite{alpha}, with further improvements expected from new low energy R data \footnote{Improved data for the critical 2-5 GeV region is already available from BES, with an eventual factor of two improvement in precision expected in this region. \cite{DPF} }. To fully exploit higher precision in $\alpha(M_Z^2)$ will also require the expected FNAL Run II improvements in $\delta m_{top}$, from the present 5 GeV to below 3 GeV. \begin{figure} [h] \begin{minipage}{5.8cm} \leavevmode\centering \epsfxsize=5.8cm \epsfbox{mhiggs2.eps} \caption{Selected MSM fits.} \end{minipage} \hfill \begin{minipage}{5.8cm} \leavevmode\centering \epsfxsize=5.8cm \epsfbox{st4ju_doe2.eps} \caption{Global S,T fit.} \end{minipage} \end{figure} A more general approach employs a fit to the S,T,U parameters \cite{ST}, which encompass a broad class of models dominated by oblique radiative effects, including supersymmetric models. We perform a global fit to all the world's electroweak data, including $sin^2\theta_W^{eff}$, $M_W$, Z-width and leptonic BRs, DIS-$\nu$ scattering, and atomic parity violation, but excluding the heavy quark results from LEP and SLD (as these may be showing significant vertex corrections). Figure 2 shows the 68\% and 90\% fit ellipses and the contributions from the three most precise inputs to the fit. The MSM allows the banana-shaped shaped region, whose size is limited by the present FNAL top mass errors, and the LEP II direct Higgs search bounds (a value of 98 GeV for the combined result is used here). Also shown are a collection of points sampled from the 5-parameter space of the Minimal Supersymetric Model (MSSM). It is evident how light Higgs masses, and hence the MSSM, are presently favored (in particular by $sin^2\theta_W^{eff}$ and $M_W$). It is also intriguiging how $sin^2\theta_W^{eff}$ has begun to place limits on MSSM parameters, and with improved precision could play a role in untangling ambiguous Higgs observations at the LHC. \section*{References}
\section{Introduction} The Standard Model (SM) has accomplished a great success to describe enormous phenomena of elementary particles. Most of the experimental data have shown quite a good quantitative agreement with the SM predictions for various observables. Precision measurements of the $Z$--pole observables at LEP and SLD have provided highly accurate tests on the Standard Model \cite{sm1,sm2,sm3}. After the controversies about the discrepancy of $R_b$ ($\equiv \Gamma(Z \to b \bar{b}) /\Gamma(Z \to \mbox{hadrons})$) are resolved along with the improvement of the experiment, no evidences of the new physics signal from colliders are reported yet. The new physics effects are thought to be as large as the loop effects of the SM at most. We are still looking for a hint of the discrepancy from the SM predictions in the list of the LEP and SLC data. The forward--backward asymmetry of $Z \to b \bar{b}$ decay, $A_{FB}^b$, may be a clue of such discrepancies as it shows $-2 \sigma$ deviation from the SM prediction. If we consider the left-right forward-backward asymmetry $A_b$, which is directly measured by SLD and is related to $A_{FB}^b$, the discrepancy is larger. The combined fit for the LEP and SLD measurements gives $A_b = 0.881 \pm 0.018$ which is $3 \sigma$ away from the SM prediction \cite{sm3}. Taking this deviation to be serious, anomalous couplings are required in the $Z b \bar{b} $ vertex. At the same time we demand that the partial decay width $\Gamma_b$ (alternatively $R_b$) should be kept within the experimental bound with these anomalous couplings. The possibility of anomalous couplings are discussed in many literatures. Field presented an detailed analyses on the asymmetries and couplings in model independent way \cite{field}. Chang and Ma suggested a vectorlike heavy quark model to explain $A_{FB}^b$ and $R_b$ data \cite{chang}. In this letter, we attempt to extract the anomalous couplings in $Z b \bar{b}$ vertex which explain $A_{FB}^b$ data of LEP and SLD together with $R_b$ through an extended scheme of electroweak precision test. The remarkable feature is that the anomalous right-handed coupling is required as well as the left-handed one since variables for asymmetries are affected by the ratio of the right-handed coupling to the left-handed one. Since the SM electroweak radiative correction is almost left-handed, the anomalous right-handed current may be a sensitive probe to new physics effects. Altarelli et al. have suggested the $\epsilon$ variables for the electroweak precision test \cite{altarelli1} and modified the analysis including $\Gamma_b$ to describe the large $m_t^2$ dependences of electroweak radiative corrections to $Z b \bar{b}$ vertex \cite{altarelli2}. However it is still dissatisfactory to describe the general couplings. Here we suggest an extension including $A_{FB}^b$ data to extract the general type of the anomalous couplings of $Z b \bar{b}$ vertex. We apply our analysis to the model in the presence of the nonuniversal contact interactions. The nonuniversal interaction acting on the third generation can be an attractive candidate for the new physics \cite{hill1,zhang,lee,hill2,hill3}, since we favor that the SM predictions for other flavours should not be much disrupted by the new physics effects. As another realization of the new physics effects, we consider the left-right model (LR model) in the general scheme. In the framework of LR model based on the extended electroweak gauge group SU(2)$_L \times$SU(2)$_R \times$U(1), the anomalous right-handed couplings are naturally introduced as well as the left-handed ones as is desirable. In this model the anomalous couplings in $b$ sector are directly related to those in the lepton sector and should be strictly constrained. For given bounds of the LR model parameter set, both of the $A_{FB}^b$ and the $R_b$ data can be shown to be accommodated. This paper is organized as follows: We extract the anomalous current interactions in $Z b \bar{b}$ vertex in terms of the model independent parameterization of the basic observables including $A_{FB}^b$ in Section II. Our analysis in terms of new $\epsilon$ variables are applied to the minimal nonuniversal contact term that is $d>4$ and its effect on the $Z b \bar{b}$ vertex in Section III. The model with SU(2)$_L \times$SU(2)$_R \times$U(1) gauge group is considered in Section IV. We present constrained parameter set $(\epsilon_b,\epsilon'_b)$ and show that both of the $A_{FB}^b$ and the $R_b$ data can be accommodated. Finally we conclude in Section V. \section{Model independent parameterization} Provided that we allow the additional contribution of the new physics to the $Z b \bar{b}$ vertex, we write the most general amplitude for $Z \to b \bar{b}$ decay for the model--independent analysis as \begin{equation} M(Z \to b \bar{b}) = \frac{g}{2 \cos \theta_W} \epsilon^\mu \bar{u} \big( \gamma_\mu (g_{bV}^0 - g_{bA}^0 \gamma_5) + \Delta^L_b \gamma_\mu P_L + \Delta^R_b \gamma_\mu P_R \big) u~, \end{equation} where $\epsilon_\mu$ is the polarization vector for $Z$ boson and $g_{bV}^0$ ($g_{bA}^0$) are tree level vector (axial--vector) couplings of $b$ quark pair to $Z$ boson given by: \begin{equation} g_{bV}^0 = -\frac{1}{2} + \frac{2}{3} s_0^2~,~~~~~ g_{bA}^0 = -\frac{1}{2}~, \end{equation} with the Weinberg angle at tree level, $s_0^2$, satisfying $s_0^2 c_0^2 = \pi \alpha(m_{_Z}) /\sqrt{2} G_F m_{_Z}^2$. In the SM, $ \Delta^L_b = \Delta^{SM}_b (m_t^2)$ and $ \Delta^R_b=0$ where the leading contribution of $\Delta^{SM}_b(m_t^2)$ in the large $m_t$ limit is given by \cite{akhundov} \begin{equation} \Delta^{SM}_b(m_t^2) = \frac{\alpha}{4 \pi \sin^2 \theta_W} | V_{tb} |^2 \frac{1}{2} \left( \frac{m_t^2}{m_{_W}^2} + \left( \frac{8}{3} + \frac{1}{6 \cos^2 \theta_W} \right) \log \frac{m_t^2}{m_{_W}^2} \right)~, \end{equation} which arises from the top quark exchange diagrams shown in Fig. 1. \begin{figure}[t] \begin{center} \epsfig{file=fig1.eps,width=10cm} \caption{ Electroweak radiative corrections to $Z b \bar{b}$ vertex. } \end{center} \vskip 0.5cm \end{figure} We present the analysis in terms of the precision variables in order to incorporate the SM corrections. It is helpful to introduce the precision variables for the study of the new physics effects on the electroweak data because the additional contributions of new physics are expected to be comparable with the loop contributions of the SM. The $\epsilon$ analysis suggested by Altarelli et al. \cite{altarelli1,altarelli2} provides a model independent way to analyze the electroweak precision data. In this scheme, the electroweak radiative corrections containing whole $m_t$ and $m_H$ dependencies are parametrized into the parameters $\epsilon$'s and thus the $\epsilon$'s can be extracted from the data without specifying $m_t$ and $m_H$. The universal correction terms to the electroweak form factors $\Delta \rho$ and $\Delta k$ are defined from the vector and axial vector couplings of lepton pairs to $Z$ boson as Eq. (4) of Ref. \cite{altarelli2} and extracted from the inclusive partial decay width $\Gamma_l$ and the forward-backward asymmetry $A_{FB}^l$. Thus we present the amplitude in terms of the vector and axial-vector couplings instead of the left- and right-handed couplings in Eq. (1). Meanwhile the SM correction (3) is left-handed and so we introduce the additional terms in the left- and right-handed basis. Another correction term $\Delta r_W$ is obtained from the mass ratio $m_{_W}/m_{_Z}$ by the Eq. (1) of Ref. \cite{altarelli2}. The $\epsilon$ parameters are defined by the linear combinations of correction terms to avoid the new physics effects being masked by the large $m_t^2$ corrections in $\epsilon_2$ and $\epsilon_3$. We note that $\Delta r_W$ is irrelevant for our analysis and affects only on $\epsilon_2$. Hence we lay aside $\epsilon_2$ in our analysis of this letter. The parameter $\epsilon_b$ is introduced to measure the additional contribution to the $Z b \bar{b}$ vertex due to the large $m_t$-dependent corrections in the SM. Since the electroweak radiative correction of the SM shown in Eq. (3) is left-handed in the large $m_t$ limit, Altarelli et al. have defined $\epsilon_b$ through the effective couplings $g_{bA}$ and $g_{bV}$ in the following manner: \begin{eqnarray} g_{bA} &=& -\frac{1}{2} (1+\frac{1}{2} \Delta \rho) (1+\epsilon_b)~, \nonumber \\ x_b &\equiv& \frac{g_{bV}}{g_{bA}} = \frac{ 1-\frac{4}{3} \sin^2 \theta^l_{eff} + \epsilon_b} {1+\epsilon_b}~. \end{eqnarray} of which asymptotic contribution is given by $\epsilon_b \approx -G_F m_t^2/4 \pi^2 \sqrt{2}$. In this expression, $ \sin^2 \theta^l_{eff}$ denotes the effective Weinberg angle including the SM loop corrections to the lepton sector. However $\epsilon_b$ in Eq. (4) cannot be the most general deviations from the Standard Model of $Z b \bar{b}$ vertex and is not suitable for incorporating new physics effects. We introduce another parameter to describe the additional right--handed current interaction effects. Firstly we define $\Delta \rho_b$ by a deviation from the axial coupling of lepton sector: \begin{equation} g_{bA} = g_{A} (1+\Delta \rho_b) = -\frac{1}{2} \left( 1+\frac{1}{2} \Delta \rho \right) (1+\Delta \rho_b) ~. \end{equation} Next we introduce another correction term $\Delta k_b$ analogous to the lepton sector as \begin{equation} \sin^2 \theta^b_{eff} = \sin^2 \theta^l_{eff} (1+\Delta k_b) = s_0^2 (1+\Delta k) (1+\Delta k_b)~, \end{equation} which leads to \begin{equation} x_b \equiv \frac{g_{bV}}{g_{bA}} = 1-\frac{4}{3} \sin^2 \theta^b_{eff} = \frac{1-\frac{4}{3} \sin^2 \theta^l_{eff} - \Delta k_b} {1 - \Delta k_b}~. \end{equation} The correction terms $\Delta \rho_b$ and $\Delta k_b$ are defined in accord with those of lepton sector. To avoid being masked by the electroweak radiative corrections of the SM, we define the new epsilon parameters by the relations \begin{equation} \epsilon_b \equiv \Delta \rho_b,~~~~~~~~ \epsilon'_b \equiv \frac{2}{3} s_0^2 (\Delta \rho_b + \Delta k_b), \end{equation} with canceling $\Delta^{SM}_b(m_t^2)$ in $\epsilon'_b$. In the SM limit, $ \Delta \rho_b =-\Delta k_b = \Delta_b^{SM}(m_t^2) $ and consequently $\epsilon'_b = 0$. Hence $\epsilon_b$ goes to the original definition in the SM limit while $\epsilon'_b$ purely measures the anomalous right-handed current interaction and goes to 0 in the SM limit. The parameters $\epsilon_b$ and $\epsilon'_b$ are extracted from the observables of the inclusive decay width $\Gamma_b$ and forward-backward asymmetry of $b \bar{b}$ production $A_{FB}^b$. In consequence, the four parameters $\epsilon_1$, $\epsilon_3$, $\epsilon_b$, and $\epsilon'_b$ are set to be one to one correspondent to the observables $\Gamma_l$, $A_{FB}^l$, $\Gamma_b$, and $A_{FB}^b$. The quadratic $m_t$ dependences of the SM electroweak radiative corrections appear in $\epsilon_1$ and $\epsilon_b$ while the $m_t$ dependence of $\epsilon_3$ is logarithmic. The parameter $\epsilon'_b$ is identical to the anomalous right-handed current interactions in $Z b \bar{b}$ vertex and expected to be sensitive to the new physics. \begin{figure}[t] \centering \epsfig{file=fig2.eps,height=12cm,width=13cm} \caption{ Available region in $\epsilon_b$ - $\epsilon'_b$ plain obtained from experiment. } \end{figure} We have the modification of the linearized relations between the observables $\Gamma_b$, $A_{FB}^b$ and the epsilon parameters as \begin{eqnarray} \Gamma_b &=& \Gamma_b |_B (1 + 1.42 \epsilon_1 - 0.54 \epsilon_3 +2.29 \epsilon_b - 1.89 \epsilon'_b) \nonumber \\ A_{FB}^b &=& A_{FB}^b |_B (1 + 17.5 \epsilon_1 - 22.75 \epsilon_3 +0.157 \epsilon_b - 1.02 \epsilon'_b) \end{eqnarray} while the relations of $\Gamma_l$ and $A_{FB}^l$ remains intact as Eq. (123) of Ref. \cite{altarelli2}. $\Gamma_b |_B$ and $ A_{FB}^b |_B $ are the Born approximation values which are defined by the tree level results including pure QED and pure QCD corrections and consequently depend upon the values of $\alpha_s(m_{_Z}^2)$ and $\alpha(m_{_Z}^2)$. The QCD corrections to the forward-backward asymmetries of $b$ quark pair in $Z$ decays are given in the Ref. \cite{abbaneo}. We obtain $ \Gamma_b |_B = 379.8$ MeV, and $A_{FB}^b |_B = 0.1041$ with the values $\alpha_s(m_{_Z}^2) = 0.119$ and $\alpha(m_{_Z}^2) =1/128.90$. We show the allowed region of $\epsilon_b$ and $\epsilon'_b$ at 95 $\%$ C.L. from the recent LEP$+$SLD data in Fig. 2. The values of $m_t = 175$ GeV and $m_H=100$ GeV are used. We obtain \begin{eqnarray} \epsilon_b &=& (1.8 - 5.4) \times 10^{-2}, \nonumber \\ \epsilon'_b &=& (2.7 - 7.0) \times 10^{-2}. \end{eqnarray} \section{Nonuniversal contact interactions} Models with nonuniversal contact interactions are mainly motivated by the idea that mass of the top quark is of order of the weak scale and so the top quark could be responsible for the electroweak symmetry breaking. In general we have several contact terms which are $d>4$ at a high energy scale in those models. We find a general list of contact terms in Refs. \cite{hill2,hill3}. As a minimal contents of the model, left-handed SU(2) doublet for the third generation and the right handed singlet $t_R$ are coupled in a new gauge interaction. The relevant term of the effective lagrangian is written by \begin{equation} L_{eff} = - \frac{1}{\Lambda} \bar{b} \gamma_{\mu} b \bar{t} \gamma_{\mu} (f_V - f_A \gamma_5) t + ...~, \end{equation} where $f_V$ and $f_A$ are model parameters and $\Lambda$ is the new physics scale. Following the Ref. \cite{hill1,hill3} for normalizations, we define $f_A \sim 4\pi(0.11)$. The effective corrections to $Z \to b \bar{b}$ decay as represented in Eq. (1) are generated via the top quark loops given by \begin{equation} \Delta_b^{L,R} = \frac{N_c}{4 \pi^2} f_A \frac{m_t^2}{\Lambda^2} \ln \frac{\Lambda^2}{m_t^2}~, \end{equation} where $N_c = 3$. Since $ \Delta_b^{L} = \Delta_b^{R}$, no additional contributions to $\epsilon_b $ and $\epsilon'_b = \Delta_b^R$. From the Fig. 2, we find that $\epsilon'_b = (0 \sim -5) \times 10^{-3}$ when $\epsilon_b = \epsilon_b^{SM}$, which yields the new physics scale $\Lambda > 1.7$ TeV. \begin{figure}[t] \begin{center} \epsfig{file=fig3.eps,width=7cm} \caption{ Vertex corrections to $Z \to b \bar{b}$ decay with the contact term. } \end{center} \vskip 1cm \end{figure} \section{SU(2)$_L \times$SU(2)$_R \times$U(1) model} If indeed the anomalous right-handed current exists in $Z b \bar{b}$ vertex, it demands to find an origin of such an anomaly as a next task. Here we consider the left-right model (LR model) based on the extended gauge group SU(2)$_L \times$SU(2)$_R \times$ U(1) as a possible candidate. We assume a general model without imposing the parity on th lagrangian where the value of $g_R$ need not be equal to $g_L$. \begin{figure}[t] \epsfig{file=fig4.eps,height=9cm,width=13cm} \caption{ Allowed parameter sets of $(\epsilon_b,\epsilon'_b)$ in LR model } \vskip 0.5cm \end{figure} In this model we have two kinds of additional contributions to the neutral current sector with right-handed current, one is the interaction to extra neutral gauge boson $Z'$ and the other is the interaction to the ordinary $Z$ boson suppressed by the mixing angle $\xi$. Since the $e^+ e^-$ collisions of LEP experiment arise at the $Z$--peak energy, the latter is relevant for our analysis, which is given by \begin{eqnarray} L_{NC} &=& \frac{e}{s_W c_W} Z^\mu \bar{b} \gamma_\mu \left[ \left( T_3^L - s_W^2 Q + \xi t_R s_W ( T_3^L-Q) \right) P_L \right. \nonumber \\ &&~~~~~~~~+ \left. \left( - s_W^2 Q + \xi (s_W (t_R+\frac{1}{t_R}) T_3^R - t_R s_W Q) \right) P_R \right] b~, \end{eqnarray} where $t_R = \tan \theta_R$, $s_W = \sin \theta_W$, $c_W = \cos \theta_W$, and $Q$ is the electric charge. $\xi$ is the neutral mixing angle between $Z$ and $Z'$. The definitions of mixing angles follow the notation of Ref. \cite{chay}. Since we do not impose a discrete L-R symmetry, the additional parameter $\theta_R$ have to come into the model besides $\xi$ and $m_{Z'}$. The correction terms defined in Eq. (1) are expressed in terms of the model parameters: \begin{eqnarray} \Delta^L_b &=& \Delta^{SM}_b (m_t^2) - \frac{1}{6} \xi t_R s_W \nonumber \\ \Delta^R_b &=& - \frac{1}{6} \xi t_R s_W \left( 1+\frac{3}{t_R^2} \right)~. \end{eqnarray} Note that the mass of extra $Z'$ boson does not enter the analysis for LEP I data since the LEP I experiment is performed at the $Z$ peak energy. \begin{figure}[t] \epsfig{file=fig5.eps,height=11cm,width=13cm} \caption{ The LR model predictions of $R_b$ and $A_{FB}^b$ with the constrained values of $\epsilon$'s. The inner ellipse denotes the experimental data at 1-$\sigma$ level and the outer ellipse at 95 $\%$ C.L.. } \vskip 0.5cm \end{figure} When we consider the LR model, the anomalous right-handed couplings to $Z$ boson also appear in other fermion sectors. Thus we have the heavy constraints on the mixing angle $\xi$ depending on $\theta_R$ from the lepton sectors \cite{chay,langacker,barenboim}. In the analysis of the neutral sector of the general LR model in Ref. \cite{chay}, the constraints on $(\xi, \theta_R)$ are obtained from the $\epsilon_1$ and $\epsilon_3$ parameters using LEP data. With these constraints on $\xi$ and Eq. (9), $\epsilon_b$ and $\epsilon'_b$ are also heavily constrained as shown in Fig. 4. Actually extremely small region out of the ellipse in the $(\epsilon_b,\epsilon'_b)$ plain shown in Fig. 2 is consistent with the allowed $(\epsilon_1,\epsilon_3)$ values in this model. With these values of $\epsilon$'s, the predictions of $A_{FB}^b$ and $R_b$ from the Eq. (9) are presented in Fig. 5 together with the experimental data. We find that there exist parameter sets with which the recent data can be consistent with the LR model predictions at 95 $\%$ C.L.. \section{Concluding Remarks} We consider the anomalous couplings of $Z b \bar{b}$ vertex to explain the discrepancy in $A_{FB}^b$ as a manifestation of new physics effects. Since there is no right--handed electroweak radiative corrections to the $Z b \bar{b}$ vertex in the large $m_t$ limit of the SM, the anomalous right--handed currents of $Z b \bar{b}$ vertex may be a sensitive probe to the new physics beyond the SM. Among the LEP observables, $A_{FB}^b$ is one of the best window to explore the anomalous right--handed currents since it is sensitive to the ratio of the right-handed coupling to the left-handed coupling. The electroweak precision variables are used to extract the new physics effects. Any new physics effects in $Z l^+l^-$ vertex are included in the $\Delta \rho$ and $\Delta k$ and consequently in $\epsilon_1$ and $\epsilon_3$. Thus the corresponding relations of $\Gamma_l$, $A_{FB}^l$ are unchanged with the new physics effects. Including the observable $A_{FB}^b$, we introduce a new variable $\epsilon'_b$ to probe the anomalous right-handed current interactions in $Z b \bar{b}$ vertex. Hence the new physics effects in $Z b \bar{b}$ vertex are encoded in $\epsilon_b$ and $\epsilon'_b$ likewise. From the experimental data we obtain the model independent bound on $\epsilon_b$ and $\epsilon'_b$. Note that the allowed parameter set $(\epsilon_1,\epsilon_3)$ should be altered when we consider another set of observables. However we find that the change is slight here since we introduce $A_{FB}^b$ and exclude $m_{_W}$ and their error pull is of the same order. We consider the nonuniversal contact interactions and the LR model as underlying physics of such an anomalous right-handed current. The lower bound of the new physics scale at which the higher dimensional operators arise is estimated through the $\epsilon$ analysis. The LR model provides shifts of $Z l^+ l^-$ couplings as well as those of $Z b \bar{b}$ vertex and they are closely related. Thus the allowed region for $(\epsilon_b,\epsilon'_b)$ set is constrained by the $(\epsilon_1,\epsilon_3)$ constraints obtained from the precise measurements of the leptonic current interactions. The large amount of deviation in $A_{FB}^b$ prefers much shift of couplings while the data of lepton sectors are rather closer to the SM predictions. Therefore the combined analysis may present a strong constraint on the neutral mixing angles of LR model if the improved analyses of $A_{FB}^b$ would be performed. \begin{ack} This work is supported by the Korea Science and Engineering Foundation through the SRC program of Center for Theoretical Physics at Seoul National University. \end{ack}
\section*{Introduction} Crossed squares were introduced by Loday and Guin-Walery in \cite{wl}. They arose in various problems of relative algebraic K-theory. Loday later showed in \cite{loday} that these quite simple algebraic gadgets modelled all homotopy 3-types. More generally his notion of cat$^n$-group and the related crossed $n$-cubes of Ellis and Steiner were shown by Loday to model all connected $(n+1)$-types. The possibilities of calculation with these models was enhanced by the development with R.Brown of a van Kampen type theorem for these structures \cite{bl1}. A link between simplicial groups and crossed $n$-cubes was used by Porter, \cite{porter} to give an algebraic form of Loday's result and in particular to give a functor from the category of simplicial groups to that of crossed $n$-cubes realising the equivalence. In 1993, Ellis \cite{ellis2} introduced a notion of free crossed square and showed how to assign a free crossed square to a CW-complex. As there was an established notion of free simplicial group, it seemed important to investigate the extent to which the two notions of freeness are related. That was the initial motivation for this paper. The two notions were intimately related and moreover combining this with Ellis' alternative description of free crossed squares in terms of the Brown-Loday non-abelian tensor product of groups and coproducts of crossed modules, gives a new purely algebraic derivation of Brown and Loday's result describing the homotopy 3-type of the suspension of an Eilenberg-Mac Lane space. This success raises our hopes that this method of attack can yield new results in higher dimensions. \section{Preliminaries } In this paper we will concentrate on the reduced case and hence on simplicial groups rather than simplicial groupoids. This is for ease of exposition only and all the results do go through for simplicially enriched groupoids. \textbf{Notation:} If $X$ is a set, $F(X)$ will denote the free group on $X$. If $Y$ is a subset of $F(X)$, $\langle Y \rangle $ will denote the normal subgroup generated by $Y$ within $F(X)$. \subsection{Simplicial groups and groupoids} Denoting the usual category of finite ordinals by $\Delta,$ we obtain for each $k\geq 0$, a subcategory $\Delta_{\leq k}$ determined by the objects $[j]$ of $\Delta$ with $j\leq k.$ A simplicial group is a functor from the opposite category $\Delta^{op}$ to $\mathfrak{Grp};$ a $k$-truncated simplicial group is a functor from $\Delta^{op}_{\leq k}$ to $\mathfrak{Grp}.$ We will denote the category of simplicial groups by $\mathfrak{SimpGrp}$ and the category of k-truncated simplicial groups by ${\mathfrak{Tr_kSimpGrp}}$. By a {\em k-truncation of a simplicial group}, we mean a $k$-truncated simplicial group $\mathfrak{tr_k}{\bf G}$ obtained by forgetting dimensions of order $>k$ in a simplicial group {\bf G}, that is restricting {\bf G} to $\Delta^{op}_{\leq k}$. This gives a truncation functor $ \mathfrak{tr_k}:{\mathfrak{SimpGrp}}\longrightarrow {\mathfrak{Tr_kSimpGrp}} $ which admits a right adjoint $ \mathfrak{ cosk_k}:\mathfrak{Tr_kSimpGrp}\longrightarrow \mathfrak{SimpGrp} $ called the {\em k-coskeleton functor}, and a left adjoint $ \mathfrak{sk_k}:\mathfrak{Tr_kSimpGrp}\longrightarrow \mathfrak{SimpGrp,} $ called the {\em k-skeleton functor}. For explicit constructions of these see \cite{duskin}. We will say that a simplicial group $G$ is \emph{k-skeletal} if the natural morphism $\mathfrak{sk_k}G\rightarrow G$ is an isomorphism. Recall that given a simplicial group {\bf G}, {\em the Moore complex} $(% { NG},\partial )$ {\em of} {\bf G} is the normal chain complex defined by $$ ({ NG})_n=\bigcap_{i=0}^{n-1}\mbox{\rm Ker}d_i^n $$ with $\partial _n:NG_n\rightarrow NG_{n-1}$ induced from $d_n^n$ by restriction. There is an alternative form of Moore complex given by the convention of taking $$\bigcap^n_{i=1} \mbox{\rm Ker}d_i^n $$ and using $d_0$ instead of $d_n$ as the boundary. One convention is used by Curtis \cite{curtis} (the $d_0$ convention) and the other by May \cite{may} (the $d_n$ convention). They lead to equivalent theories. The {\em n$^{th}$ homotopy group} $\pi _n$({\bf G}) of {\bf G} is the $n ^{th}$ homology of the Moore complex of {\bf G}, i.e. $$ \begin{array}{rcl} \pi _n({\bf G}) & \cong & H_n( {NG},\partial ) \\ & = & \bigcap\limits_{i=0}^n\mbox{\rm Ker}% d_i^n/d_{n+1}^{n+1}(\bigcap\limits_{i=0}^n\mbox{\rm Ker}d_i^{n+1}). \end{array} $$ We say that the Moore complex {\bf NG} of a simplicial group is of {\em length} $k$ if $NG_n=1$ for all $n\geq k+1$, so that a Moore complex of length $k$ is also of length $l$ for $l\geq k.$ For example, if ${\bf G}$ has Moore complex of length 1, then $(NG_1,NG_0,\partial_1)$ is a crossed module and conversely. If $NG$ is of length 2, the corresponding Moore complex gives a 2-crossed module in the sense of Conduch\'e, \cite{conduche}, cf. the companion paper to this, \cite{mp3} \subsection{Free Simplicial Groups} Recall from \cite{curtis} and \cite{kan2} the definitions of free simplicial group and of a $CW-basis$ for a free simplicial group. {\bf Definition} A simplicial group {\bf F} is called \emph{free} if\\ (a)\qquad $F_n$ is a free group with a given basis, for every integer $n\geq 0,$\\ (b)\qquad The bases are stable under all degeneracy operators, i.e., for every pair of integers $(i,n)$ with $0\leq i\leq n$ and every basic generator $x\in F_n$ the element $s_i(x)$ is a basic generator of $F_{n+1}.$ {\bf Definition} Let ${\bf F}$ be a free simplicial group (as above). A subset $\mathfrak{F}\subset {\bf F}$ will be called a $CW-basis$ of ${\bf F}$ if \\ (a)\qquad $\mathfrak{F_n} = \mathfrak{F}\cap F_n$ freely generates $F_n$ for all $n\geq 0,$\\ (b)\qquad $\mathfrak{F}$ is closed under degeneracies, i.e. $x\in \mathfrak{F_n}$ implies $s_i(x)\in \mathfrak{F_{n+1}}$ for all $0\leq i\leq n,$\\ (c)\qquad if $x\in\mathfrak{F_n}$ is non-degenerate, then $d_i(x) = e_{n-1},$ the identity element of $F_n$, for all $0\leq i< n$. As explained earlier, we have restricted attention so far to simplicial groups and hence to connected homotopy types. This is traditional but a bit unnatural as all the results and definitions so far extend with little or no trouble to simplicial groupoids in the sense of Dwyer and Kan \cite{D&K} and hence to non-connected homotopy types. It should be noted that such simplicial groupoids have a fixed and constant simplicial set of objects and so are not merely simplicial objects in the category of groupoids. In this context if $\mathbf{G}$ is a simplicial groupoid with set of objects $O$, the natural form of the Moore complex $\mathbf{NG}$ is given by the same formula as in the reduced case, interpreting Ker$ d^n_i$ as being the subgroupoid of elements in $G_n$ whose $i^{th}$ face is an identity of $G_{n-1}$. Of course if $n \geq 1$, the resulting $NG_n$ is a disjoint union of groups, so $\mathbf{NG}$ is a disjoint union of the Moore complexes of the vertex simplicial groups of $\mathbf{G}$ together with the groupoid $G_0$ providing elements that allow conjugation between (some of) these vertex complexes (cf. Ehlers and Porter \cite{ep}). Crossed modules of, or over, groupoids are well known from the work of Brown and Higgins. The only changes from the definition for groups (cf. \cite{loday}) is that one has to handle the conjugation operation slightly more carefully: A \emph{crossed module} is a morphism of groupoids $\partial : M\longrightarrow N$ where $N$ is a groupoid with object set $O$ say and $M$ is a family of groups, $M = \{M(a) : a \in O\}$, together with an action of $N$ on $M$ satisfying (i) if $m \in M(a)$ and $n \in N(a,b) $ for $a,b,\in O$, the result of $n$ acting on $m$ is ${}^nm \in M(b)$; (ii) $\partial({}^nm)=n\partial(m)n^{-1}$ and (iii) ${}^{\partial(m)}{m'}=m{m'}m^{-1}$ for all $m,{m'}\in M$, $\ n\in N.$ For the weaker notion in which condition (iii) is not required, the models are called \emph{precrossed modules}. The definition of a CW-basis likewise generalises with each $\mathfrak{F}$ a subgraph of the corresponding free simplicial groupoid. \section{Crossed Squares and Simplicial Groups} Although we will be mainly concerned with crossed squares in this paper, many of the arguments either clearly apply or would seem to apply in the more general case of crossed $n$-cubes and $n$-cube complexes. We therefore give some background in this more general setting. Again although we give the definitions and results for groups, the adaptation to handle groupoids over a fixed base is routine. The following definition is due to Ellis and Steiner \cite{es}. Let $<n>$ denote the set $\{1,...,n\}.$ {\bf Definition} A \emph{crossed $n$-cube} of groups is a family $ \{ \mathfrak{M}_A : A\subseteq <n>\}$ of groups, together with homomorphisms $\mu _i:\mathfrak{M}_A\longrightarrow \mathfrak{M}_{A\setminus\{i\}}$ for $i\in <n>$ and functions $$h:\mathfrak{M}_A\times \mathfrak{M}_B\longrightarrow \mathfrak{M}_{A\cup B}$$for $A,B\subseteq <n>,$ such that if ${}^{a}b$ denotes $h(a,b)b$ for $a \in \mathfrak{M}_{A}$ and $b \in \mathfrak{M}_{B}$ with $A\subseteq B,$ then for all $a,{a'} \in \mathfrak{M}_{A}$ and $b,{b'} \in \mathfrak{M}_{B}, c \in \mathfrak{M}_{C}$ and $i,j \in <n>,$ the following hold: $$ \begin{array}{ll} 1) & \mu _ia = a\ \quad \text{{\rm if}}\ i\not \in A, \\ 2) & \mu _i\mu _ja = \mu _j\mu _ia, \\ 3) & \mu _ih(a, ~b) = h(\mu _ia, ~\mu _ib), \\ 4) & h(a, ~b) = h(\mu _ia, ~b)=h(a, ~\mu _ib) \hfill \text{{\rm if}}\ i\in A\cap B, \\ 5) & h(a, ~a^{\prime }) =\lbrack a,{~}a^{\prime } \rbrack, \\ 6) & h(a, ~b) = h{(b, ~a)}^{-1}, \\ 7) & h(a, ~b)=1 \hfill\text{if $a = 1$ ~~\text{or}~~ $b = 1,$}\\ 8) & h(aa^{\prime }, ~b)={}^{a} h(a^{\prime }, ~b)h(a, ~b), \\ 9) & h(a,~bb^{\prime })= h(a,~b){~}{}^{b}h(a, ~b^{\prime }), \\ 10) & {}^{a}h(b, ~c)= h({}^{a}b,{~}{}^{a}c) \hfill \text{{\rm if }} A \subseteq B\cap C,\\ 11)&{}^{a}h(h(a^{-1}, ~b), ~c)~{}^{c}h(h(c^{-1}, ~a), ~b)~{}^{b}h(h(b^{-1}, ~c), ~a) = 1. \\ \end{array} $$ {\em A morphism of crossed n-cubes} is defined in the obvious way: It is a family of group homomorphisms, for $A\subseteq <n>,$ $ f_A:\mathfrak{M}_A\longrightarrow \mathfrak{M}_{A}^\prime $ commuting with the $\mu _i$'s and $h$'s. We thus obtain a category of crossed $n$-cubes which will be denoted by $\mathfrak{Crs^n},$ cf. Ellis and Steiner \cite{es}. Again there is an obvious variant of this definition for groupoids over a fixed set of objects, $O$. {\bf Remark:} Crossed squares, that is the case $n = 2$, were introduced by Loday and Guin-Walery, \cite{wl}, but with an apparently different definition. The two notions are however equivalent. {\noindent{\bf Example 1:}} For $n=1,$ a crossed 1-cube is the same as a crossed module. For $n=2,$ one has a crossed 2-cube is a crossed square: $$ \diagram \mathfrak{M}_{<2>} \dto_{\mu_1} \rto^{\mu_2} & \mathfrak{M}_{\{1\}} \dto^{\mu_1} \\ \mathfrak{M}_{\{2\}} \rto_{\mu_2} & \mathfrak{M}_{\emptyset}. \enddiagram $$ Each $\mu _i$ is a crossed module, as is $\mu _1\mu _2$. The $h$-functions give actions and a function $$ h:\mathfrak{M}_{\{1\}}\times \mathfrak{M}_{\{2\}}\longrightarrow \mathfrak{M}_{<2>}. $$ The maps $\mu _2$ also define a map of crossed modules from $(\mathfrak{M}_{<2>},\mathfrak{M}_{\{2\}} ,\mu_1)$ to $(\mathfrak{M}_{<1>},\mathfrak{M}_{\emptyset},\mu_1)$. In fact a crossed square can be thought of as a crossed module in the category of crossed modules. {\noindent\bf Example 2:} Let ${N_1}, {N_2}$ be normal subgroups of a group $G$. The commutative square diagram of inclusions; $$ \diagram {N_1} \cap {N_2}\dto_{~~~~~~~~~inc.} \rto^{inc.} & {N_2} \dto^{inc.} \\ {N_1} \rto_{inc.} & G \enddiagram $$ naturally comes together with actions of ${G}$ on ${N_1},{N_2}$ and ${N_1}\cap {N_2}$ given by conjugation and functions $$ \begin{array}{cccc} h: & {N_A}\times {N_B}& \longrightarrow & {N_A}\cap {N_B} = N_{A\cup B} \\ & (n_1,n_2) & \longmapsto & \lbrack n_1,~n_2\rbrack. \end{array} $$ That this is a crossed square is easily checked. The following proposition is noted by the second author in \cite{porter}. \begin{prop}\label{alti} \cite{porter} Let ${\bf G}$ be a simplicial group with simplicial normal subgroups ${\bf N_1}$ and ${\bf N_2}.$ Then the square% $$ \diagram {\bf N_1}\cap {\bf N_2}\dto \rto &{\bf N_2} \dto \\ {\bf N_1} \rto &{\bf G} \enddiagram $$ induces a crossed square $$ \diagram \pi_0({\bf N_1}\cap {\bf N_2}) \dto \rto &\pi_0({\bf N_2}) \dto \\ \pi_0({\bf N_1})\rto &\pi_0({\bf G}). \enddiagram $$ \end{prop} \begin{pf} The $h$-function $$ h:\pi _0({\bf {N}_1})\times \pi _0({\bf {N}_2})\longrightarrow \pi _0({\bf {N}_1}\cap {\bf N_2}) $$ is given by $$ h({\overline{n_1}},~{\overline{n_2}})= {\overline{[n_1,~n_2]}} $$ for all $ {\overline{n_1}} \in \pi _0({\bf {N}_1}),\,{\overline{n_2}} \in \pi _0({\bf {N}_2}).$ It is then simple, cf. \cite{porter}, to see that the second diagram above is a crossed square. {~}\end{pf}\\ In fact up to isomorphism all crossed squares arise in this way, cf. \cite{loday} and \cite{porter}. {\bf Example 3:}\label{z} Let {\bf G} be a simplicial group. Let $\mathfrak{{M}}({\bf G},2)$ denote the following diagram $$ \diagram NG_2/\partial_3NG_3 \dto_{~~~~~\partial_2 '} \rto^{\qquad \partial_2} & NG_1\dto^{\mu} \\ \overline{NG_1} \rto_{~~~\mu'} & G_1 \enddiagram $$ Then this is the underlying square of a crossed square. The extra structure is given as follows: $NG_1=${\rm Ker}$d_0^1$ and $\overline{NG}_1=${\rm Ker}$d_1^1$. Since $G_1$ acts on $NG_2/\partial _3NG_3,\ \overline{NG}_1$ and $NG_1,$ there are actions of $\overline{NG}_1$ on $NG_2/\partial _3NG_3$ and $NG_1$ via ${\mu'},$ and $NG_1$ acts on $NG_2/\partial _3NG_3$ and $\overline{NG} _1$ via $\mu.$ Both $\mu $ and ${\mu'}$ are inclusions, and all actions are given by conjugation. The $h$-map is $$ \begin{array}{ccl} NG_1\times \overline{NG}_1 & \longrightarrow & NG_2/\partial _3NG_3 \\ (x,\overline{y}) & \longmapsto & h(x,~y)= \lbrack s_1x,~s_1ys_0{y}^{-1} \rbrack \partial_3NG_3. \end{array} $$ Here $x$ and $y$ are in $NG_1$ as there is a bijection between $NG_1$ and $\overline{NG}_1.$ We leave the verification of the axioms of a crossed square to the reader.\\ This example is clearly functorial and we denote by $$ \diagram \mathfrak{M}( - ,2)~:~\mathfrak{SimpGrp} \rto & \mathfrak{Crs^2}, \enddiagram $$ the resulting functor. This is the case $n = 2$ of a general construction of a crossed $n$-cube from a simplicial group given by the second author in \cite{porter} based on some ideas of Loday. {\noindent{\bf Examples 2 and 3 revisited:}} Let $G$ be a group with normal subgroups ${N}_1, \ldots ,{N}_n$ of $G$. Let $$ \begin{array}{ccc} \mathfrak{M}_A=\bigcap \{{N}_i:i\in A\} & \text{and} & \mathfrak{M}_\emptyset =${G}$ \end{array} $$ with $A\subseteq <n>.$ For $i\in <n>,$ $\mathfrak{M}_A$ is a normal subgroup of $\mathfrak{M}_{A-\{i\}}$. Define $$ \mu _i:\mathfrak{M}_A\longrightarrow \mathfrak{M}_{A-\{i\}} $$ to be the inclusion. If $A,B\subseteq <n>$, then $\mathfrak{M}_{A\cup B}=\mathfrak{M}_A\cap \mathfrak{M}_B,$ let $$ \begin{array}{cccl} h: & \mathfrak{M}_A\times\mathfrak{M}_B & \longrightarrow & \mathfrak{M}_{A\cup B} \\ & (a,b) & \longmapsto & [{~}a,{~}b{~}] \end{array} $$ as $[\mathfrak{M}_A, \mathfrak{M}_B] \subseteq \mathfrak{M}_A\cap \mathfrak{M}_B,$ where $a\in \mathfrak{M}_A,\ b\in \mathfrak{M}_B.$ Then $$ \{\mathfrak{M}_A:\ A\subseteq <n>,\ \mu _i,\ h\} $$ is a crossed $n$-cube, called the {\em inclusion crossed n-cube} given by the normal $n$-ad of groups $(G;\ {N}_1, \ldots, {N}_n).$ \begin{prop}\label{y} Let $({\bf G};\ {N}_1, \ldots, {N}_n)$ be a simplicial normal $n$-ad of subgroups of groups and define for $A\subseteq <n>$% $$ \mathfrak{M}_A=\pi _0(\bigcap\limits_{i\in A}{N}_i) $$ with homomorphisms $\mu _i:\mathfrak{M}_A\longrightarrow \mathfrak{M}_{A-\{i\}}$ and h-maps induced by the corresponding maps in the simplicial inclusion crossed $n$-cube, constructed by applying the previous example to each level. Then $\{\mathfrak{M}_A:\ A\subseteq <n>,\ \mu _i,\ h\}$ is a crossed $n$-cube. \end{prop} \begin{pf} See \cite{porter}. \end{pf}\\ This describes a functor, \cite{porter}, from the category of simplicial groups to that of crossed $n$-cubes of groups. \begin{thm}\label{t1} If {\bf G} is a simplicial group, then the crossed $n$-cube $\mathfrak{M}${\rm (}% {\bf G},$n${\rm )} is determined by: (i) for $A\subseteq <n>,$% $$ \mathfrak{M}({\bf G},n)_A=\frac{\bigcap_{j\in A}\text{{\rm Ker}} d_{j-1}^n}{ d_{n+1}^{n+1}(\text{{\rm Ker}} d_0^{n+1}\cap \{\bigcap_{j\in A}\text{{\rm Ker} } d_j^{n+1}\})}; $$ (ii) the inclusion $$ \bigcap_{j\in A}\text{{\rm Ker}} d_{j-1}^n\longrightarrow \bigcap_{j\in A-\{i\}}\text{{\rm Ker}} d_{j-1}^n $$ induces the morphism $$ \mu _i:\mathfrak{M}({\bf G},n)_A\longrightarrow \mathfrak{M}({\bf G},n)_{A-\{i\}}; $$ (iii) the functions, for $A,B\subseteq <n>,$ $$ h:\mathfrak{M}({\bf G},n)_A\times \mathfrak{M}({\bf G},n)_B\longrightarrow \mathfrak{M}({\bf G},n)_{A\cup B} $$ are given by $$ h(\bar x,\bar y) = \overline{[x,~y]}, $$ where an element of $\mathfrak{M}({\bf G},n)_A$ is denoted by $\bar{x}$ with $x\in \bigcap_{j\in A}${\rm Ker}$d_{j-1}^n.$ \end{thm}\hfill$\Box$ Some simplification is possible, again see \cite{porter} for the details. \begin{prop}\label{w} If {\bf G} is a simplicial group, then i) for $A\subseteq <n>,\ A\neq <n>,$ $$ \mathfrak{M}({\bf G},n)_A\cong \bigcap_{i\in A}\text{{\rm Ker}}d_{i-1}^{n-1} $$ so that in particular, $\mathfrak{M}({\bf G},n)_\emptyset \cong G_{n-1}$; in every case the isomorphism is induced by $d_0,$ ii) if $A\neq <n>$ and $i\in <n>,$ $$ \mu _i:\mathfrak{M}({\bf G},n)_A\longrightarrow \mathfrak{M}({\bf G},n)_{A\setminus \{i\}} $$ is the inclusion of a normal simplicial subgroup, iii) for $j\in <n>,$ $$ \mu _j:\mathfrak{M}({\bf G},n)_{<n>}\longrightarrow \bigcap_{i\neq j}\text{{\rm Ker}}% d_i^{n+1} $$ is induced by $d_n.$ \end{prop}\hfill$\Box$ Expanding this data out for low values of $n$ gives:\\ \noindent 1) For $n=0$, $$ \begin{array}{rcl} \mathfrak{M}({\bf G},0) & = & G_0/d_1( \text{{\rm Ker}}d_0,) \\ & \cong & \pi _0( {\bf G}), \\ & = & H_0(N{\bf G).} \end{array} $$ 2) For $n=1,$ $\mathfrak{M}({\bf G},1)$ is the crossed module $$ \mu_1 :\text{{\rm Ker}}d_0^1/d_2^2(NG_2)\longrightarrow G_1/d_2^2(\text{{\rm Ker}}d_0^2). $$ 3) For $n=2,$ $\mathfrak{M}({\bf G},2)$ is $$ \diagram \mbox{\rm Ker}d^{2}_{0} \cap \mbox{\rm Ker}d^{2}_{1} / d^{3}_{3} (\mbox{ \rm Ker}d^{3}_{0} \cap \mbox{\rm Ker}d^{3}_{1} \cap \mbox{\rm Ker}d^{3}_{2})\dto_{\quad\mu_1} \rto^{\qquad\quad \mu_2} & \mbox{\rm Ker}d^{2}_{0} / d^{3}_{3}(\mbox{\rm Ker}d^{3}_{0} \cap \mbox{\rm Ker}d^{3}_{1}) \dto^{\mu_1} \\ \mbox{\rm Ker}d^{2}_{1} / d^{3}_{3}(\mbox{\rm Ker }d^{3}_{0} \cap \mbox{\rm Ker}d^{3}_{2}) \rto_{~~~~~\mu_2} & G_2 / d^{3}_{3} (\mbox{\rm Ker}d^{3}_{0}). \enddiagram $$ By Proposition \ref{w}, this is isomorphic to $$ \diagram NG_2 / d^{3}_{3} (NG_3 ) \dto_{\mu_1} \rto^{\quad\mu_2} & \mbox{\rm Ker }d^{1}_{0} \dto^{\mu_1} \\ \mbox{\rm Ker }d^{1}_{1} \rto_{\mu_2} & G_1, \enddiagram $$ that is $$ \mathfrak{M}(\mathbf{G}, 2) \cong \left ( \diagram NG_2 / \partial_3 (NG_3 ) \dto \rto & \mbox{ Ker }d_{0} \dto \\ \mbox{ Ker }d_{1} \rto & G_1 \enddiagram \right) $$ is a crossed square. Here the $h$-map is $$h: \mbox{Ker}d_0^1\times\mbox{Ker}d_1^1\longrightarrow NG_2/d_3^3(NG_3)$$ given by $h(x,y) = [s_1x, ~s_1ys_0y^{-1}]~\partial_3NG_3$, as before. Note if we consider the above crossed square as a vertical morphism of crossed modules, we can take its kernel and cokernel within the category of crossed modules. In the above, the morphisms in the top left hand corner are induced from $d_2$ so $$ \mbox{Ker}\left ( \mu_1 : \frac{NG_2}{\partial_3NG_3}\longrightarrow \mbox{Ker}d_1 \right) = \frac{NG_2\cap \mbox{Ker}d_2}{\partial_3NG_3} \cong \pi_2({\bf G}) $$ whilst the other map labelled $\mu_1$ is an inclusion so has trivial kernel. Hence the kernel of this morphism of crossed modules is $$ \pi_2({\bf G})\longrightarrow 1. $$ The image of $\mu_2$ is closed and normal in both the groups on the bottom line and as $\mbox{Ker}d_0 =NG_1$ with the corresponding $\mbox{Im}\mu_1$ being $d_2NG_2,$ the cokernel is $NG_1/\partial_2NG_2,$ whilst $G_1/\mbox{Ker}d_0\cong G_0,$ i.e., the cokernel of $\mu_1$ is $\mathfrak{M}({\bf G}, 1).$ In fact of course $\mu_1$ is not only a morphism of crossed modules, it is a crossed module. This means that $\pi_2({\bf G})\longrightarrow 1$ is in some sense a $\mathfrak{M}({\bf G}, 1)$-module and that $\mathfrak{M}({\bf G}, 2)$ can be thought of as a crossed extension of $\mathfrak{M}({\bf G}, 1)$ by $\pi_2({\bf G}).$ \section{Free Crossed Squares} \subsection{Definitions} G. Ellis, \cite{ellis2}, in 1993 presented the notion of a free crossed square. In this section, we recall his definition and give a construction of free crossed squares by using the second dimensional Peiffer elements and the $2$-skeleton of a `step-by-step' construction of a free simplicial group with given $CW$-basis. We firstly recall the definition of a free crossed square on a pair of functions $(f_2,f_3)$, as given by Ellis. We will call these crossed squares \emph{totally free}. Let ${\bf B_1}, \ {\bf B_2}$ and ${\bf B_3}$ be sets. Take $F({\bf B_1})$ to be the free group on ${\bf B_1}.$ Suppose given a function $f_2:{\bf B_2\longrightarrow }F({\bf B_1}).$ Let $\partial : M\longrightarrow F({\bf B_1})$ be the free pre-crossed module on $f_2.$ Using the action of $F({\bf B_1})$ on $M$ we can form the semi-direct product $M\rtimes F({\bf B_1}).$ The canonical inclusion $\mu: M \longrightarrow M\rtimes F({\bf B_1})$ given by $m\mapsto(m,1)$ allows us to consider $M$ as a normal subgroup of $M\rtimes F({\bf B_1}).$ (Recall that any normal inclusion is a crossed module with action given by conjugation.) There is a second normal subgroup of $M\rtimes F({\bf B_1})$ arising from $M,$ namely $$ N = \{(m,\partial m^{-1}): m\in M\}\subset M\rtimes F({\bf B_1}) $$ with inclusion denoted ${\mu'} : N\longrightarrow M\rtimes F({\bf B_1}).$ For $m\in M$, we let ${m'}$ denote the element $(m^{-1},~\partial m)$ in $N.$ Assume given a function $f_3: {\bf B_3}\longrightarrow M,$ whose image lies in the kernel of the homomorphism $\partial:M\longrightarrow F({\bf B_1}).$ There is then a corresponding function ${f_3'} : {\bf B_3}\longrightarrow N$ given by $y \mapsto (f_3(y), 1).$ {\bf Definition} \cite{ellis2} A crossed square, $$\xymatrix{L\ar[r]^{\partial_2}\ar[d]_{\partial_2^\prime}& M\ar[d]^\mu \\ N\ar[r]_{\mu^\prime\hspace{5mm}}& M\rtimes F({\bf B_1}),}$$ is totally free on the pair of functions $(f_2, f_3)$ if \\ (i) $(M,F({\bf B_1}), \partial)$ is the free pre-crossed module on $f_2$;\\ (ii)$\quad {\bf B_3}$ is a subset of $L$ with $f_3$ and ${f_3'}$ the restrictions of $\partial_2$ and ${\partial_2'}$ respectively; \\ (iii) for any crossed square$$\xymatrix{L^{\prime}\ar[r]^{\tau}\ar[d]_{\tau^\prime}& M\ar[d]^\mu \\ N\ar[r]_{\mu^\prime\hspace{5mm}}& M\rtimes F({\bf B_1}),}$$ and any function $\nu :{\bf B_3}\longrightarrow {L'}$ satisfying ${\tau}\nu = f_3,$ there is a unique morphism $\Phi= (\phi, 1, 1, 1)$ of crossed squares: $$ \diagram L \ddto_{\partial_2'} \drto^{\phi} \rrto^{\partial_2} && {M} \xto'[1,0]_{\mu}[2,0] \drto^{=} \\ &{{L'}} \ddto\save\go[0,0];[1,0]:(0.5,-0.15)\drop{\tau^\prime}\restore \rrto^{\qquad\tau} && {M} \ddto^{\mu} \\ N \drto_{=} \xline'[0,1]^{\mu'}[0,2]|>\tip && M \rtimes {F({\bf B_1})} \drto^{=} \\ &N \rrto^{\mu'} && M \rtimes F({\bf B_1})\\ \enddiagram $$ such that $\phi{\nu'} =\nu,$ where ${\nu'}:{\bf B_3} \longrightarrow L$ is the inclusion. We denote such a totally free crossed square by $(L,M,N,M\rtimes {F({\bf B_1})})$ omitting the structural morphisms from the notation when there is no danger of confusion. We know the free pre-crossed module on $f_2: {\bf B_2} \longrightarrow F({\bf B_1})$ is $\partial : \langle{\bf B_2}\rangle \longrightarrow F({\bf B_1})$, where $\langle{\bf B_2}\rangle$ denotes the normal closure of ${\bf B_2}$ in the free group $F({\bf B_2}\cup s_0({\bf B_1}))$, so the function $f_3 : {\bf B_3}\longrightarrow M ~~( =\langle{\bf B_2}\rangle )$ is precisely the data $({\bf B_3}, f_3)$ for 2-dimensional construction data in the simplicial context, cf. \cite{mp2}. We thus need to recall the $2$-dimensional construction for a free simplicial group. This $2$-dimensional form can be summarised by the diagram $${\bf \mathbb{F}}^{(2)}: \diagram ...{~}F(s_1s_0({\bf B_1})\cup s_0({\bf B_2}) \cup s_1({\bf B_2})\cup {\bf B_3}) \rto<0.25ex> \rto<1ex> \rto<1.75ex>^{\qquad\qquad\quad d_0 ,d_1 ,d_2 } & F(s_0({\bf B_1})\cup {\bf B_2}) \lto<0.75ex> \lto<1.50ex>^{\qquad\qquad\quad s_1,s_0} \rto<0.25ex> \rto<1ex>^{\qquad d_1, d_0} & F({\bf B_1}) \lto<0.75ex>^{\qquad s_0} \enddiagram $$ with the simplicial morphisms given as in \cite{mp2}. \subsection{Free crossed squares exist.} \begin{thm} A totally free crossed square $(L, M, N, M\rtimes {F({\bf B_1})})$ exists on the 2-dimensional construction data and is given by $\mathfrak{M}(\bf{F}^{(2)},2)$ where $\bf{F}^{(2)}$ is the 2-skeletal free simplicial group defined by the construction data. \end{thm} \begin{pf} Suppose given the 2-dimensional construction data for a free simplicial group, $\mathbf{F},$ which we will take as above as the data for a totally free crossed square. We will not assume detailed knowledge of \cite{mp2} so we start with $F({\bf B_1})$ and $f_2:{\bf B_2}\longrightarrow F({\bf B_1})$ and form $M=\langle{\bf B_2}\rangle.$ This gives $\partial_1: \langle{\bf B_2}\rangle \longrightarrow F(X_0)$ as the free pre-crossed module on $f_2.$ The semidirect product gives $$ F(s_0({\bf B_1})\cup {\bf B_2})\cong M\rtimes F({\bf B_1}) $$ and we can identify this with $\bf{F}_1^{(2)}.$ This identification also makes $$ M\cong \mbox{Ker}d_0^1 $$ for the $d_0^1$ of $\mathbf{F}^{(2)}.$ Next form $N =\{ (m,\partial m^{-1})\in M\rtimes F({\bf B_1}): m \in M\}$. As $m\in\langle{\bf B_2}\rangle$, it is a product of conjugates of elements of ${\bf B_2}$ and their inverses, so writing $m=\prod(m_{\alpha_i}) y_{\alpha_i}^{\varepsilon_i}(m_{\alpha_i})^{-1}$ for indices $\alpha_i$, and $\varepsilon_i =\pm 1$, we get $\partial m = \prod m_{\alpha_i} t_{\alpha_i}^{\varepsilon_i}m_{\alpha}^{-1}$ where $t_i =f_2(y_i)$, which is also $ d_0^1(y_i)$. Thus we can identify $N$ with $\langle \{ys_1d_0^1(y)^{-1}: y \in {\bf B_2}\}\rangle$, which is exactly Ker$d_1^1$. Now $f_3:{\bf B_3}\rightarrow \mbox{\rm Ker}\partial_1 = \mbox{\rm Ker} (\partial : NF^{(2)}_1\rightarrow NF^{(2)}_0)\subset \langle {\bf B_2}\rangle.$ We know that this allows us to construct ${\bf F}_2^{(2)}$ and hence ${\bf F}_n^{(2)}$ for ~$n\geq 3$, and in addition that taking $$L=NF^{(2)}_{2}/\partial _3(NF^{(2)}_3),$$ gives a crossed square $$ \diagram L\dto_{\partial'}\rto^{\partial} & {M}\dto^{\mu}\\ N\rto_{\mu'} & F^{(2)}_1 \enddiagram $$ which is $\mathfrak{M}({\mathbf{F}}^{(2)}, 2).$ We claim this is the totally free crossed square on the construction data. At this stage it is worth noting that there seems to be no simple adjointness statement between $\mathfrak{M}(-,2)$ and some functor that would give a quick proof of freeness. The problem is that $\mathfrak{M}(-,2)$ seems to be an adjoint only up to some sort of coherent homotopy. To avoid this difficulty we use a more combinatorial approach involving the higher dimension Peiffer elements and the explicit description of $L$. In \cite{mp1}, we analysed in general the structure of groups of boundaries such as $\partial _3 (NF_3^{(2)})$. There we showed that $ NF_3^{(2)}$ is normally generated by elements of the following forms:-\\ (i) For all $x\in NF^{(2)}_1,~y\in NF^{(2)}_2,$ $$ \begin{array}{lcl} f_{(1,0)(2)}(x , y) & = & [s_1s_0(x) , s_2(y)] [s_2(y) , s_2s_{0}(x)], \\ f_{(2,0)(1)}(x , y) & = & [s_2s_0(x) , s_1(y)] [s_1(y) , s_2s_1(x)] [s_2s_1(x) ,s_2(y)] [s_2(y) , s_2s_0(x)] ; \end{array} $$ (ii) for all $ y \in NF^{(2)}_{2} , x \in NF^{(2)}_{1},$ $$ \begin{array}{lcl} f_{(0)(2,1)}(x , y) & = & [s_0(x) , s_2s_1(y)] [s_2s_1(y) , s_1(x)] [s_2(x) , s_2s_1(y)], \end{array} $$ and (iii) for all $x , y \in NF^{(2)}_2$, $$ \begin{array}{rcl} f_{(0)(1)}(x , y) & = & [s_0(x) , s_1(y)] [s_1(y) , s_1(x)] [s_2(x) ,s_2(y)], \\ f_{(0)(2)}(x , y) & = & [s_0(x) , s_2(y)], \\ f_{(1)(2)}(x , y) & = & [s_1(x) , s_2(y)] [s_2(y) , s_2(x)]. \end{array} $$ Given our description of $NF^{(2)}$ in low dimensions, it is routine to calculate normal generators of the various groups involved here in terms of ${\bf B_1}$ and ${\bf B_2}$. We set $$Z=\{s_1(y)^{-1}s_0(y): y\in{\bf B_2}\}.$$ The above diagram can then be realised as $$ \diagram {J}\dto_{\partial_2'} \rto^{\partial_2} &{\langle{\bf B_2}\rangle} \dto^{\mu} & \\ \langle Z\rangle\rto_{{\mu'}\qquad} & \langle{\bf B_2}\rangle\rtimes F({\bf B_1}) \enddiagram .$$ Here $J$ is $(\langle s_1({\bf B_2})\cup {\bf B_3}\rangle \cap \langle Z\cup{\bf B_3}\rangle )/P_2$, $P_2$ being the second dimensional Peiffer normal subgroup, which is in fact just $\partial _3 (NF_3^{(2)})$, and which is a subgroup of $\langle s_1({\bf B_2})\cup {\bf B_3}\rangle \cap\langle Z\cup{\bf B_3}\rangle$. Given any crossed square $({L'},M, N, M\rtimes {F({\bf B_1})})$ and a function $\nu: {\bf B_3}\longrightarrow {L'},$ there then exists a unique morphism $$ \phi :(L,M,N,M\rtimes {F({\bf B_1})})\longrightarrow ({L'},M,N, M\rtimes {F({\bf B_1})}) $$ given by% $$ \phi ({y_i'}P_2)=\nu ({y_i'}) $$ such that $\phi{\nu'} =\nu.$ The existence of $\phi$ follows by using the freeness property of the group $NF_2^{(2)}$ and then restricting to $\langle s_1({\bf B_2})\cup {\bf B_3}\rangle \cap\langle Z\cup{\bf B_3}\rangle$. The normal generating elements of $P_2$ are then easily shown to have trivial image in $L'$ as that group is part of the second crossed square. Thus the diagram is the desired totally free crossed square on the 2-dimensional construction data. The crossed square properties of $(L,M,N,M\rtimes {F({\bf B_1})}$ may be easily verified or derived from the fact that this is exactly $\mathfrak{M}({\bf{F}}^{(2)}, 2)$. \end{pf} \medskip {\bf Remark:} At this stage, it is important to note that nowhere in the argument was use made of the freeness of the 1-skeleton. If $G$ is any 1-skeletal simplicial group and we form a new simplicial group $H$ by adding in a set ${\bf B_3}$ of new generators in dimension 2, so that for instance, $H_2 = G_2 * F({\bf B_3})$, then we can use $M = NG_1 = {\rm Ker }d_0^{G,1}$ as before even though it need not be free. The corresponding $N$ is then isomorphic to ${\rm Ker }d_1^{G,1}$ with the bottom right hand corner being $G_1$. The `construction data' is now replaced by data for killing some elements of $\pi_1(G)$, specified by $f_3 :{\bf B_3} \rightarrow M$. Although slightly at variance with the terminology used by Ellis, \cite{ellis2}, we felt it sensible to introduce the term ``totally free crossed square'' for the type of free crossed square constructed in the above theorem, using ``free crossed square'' for the more general situation in which $(M,G,\partial)$ and $f_3$ are specified and no requirement on $(M,G,\partial)$ to be a free precrossed module is made. \subsection{The $n$-type of the $k$-skeleton} As in the other papers in this series, we will use the `step-by-step' construction of a free simplicial group to observe the way in which the models react to the various steps of the construction. In a `step-by-step' construction of a free simplicial group, there are simplicial inclusions $$ {\bf{F}}^{(0)}\subseteq {\bf{F}}^{(1)}\subseteq {\bf{F}}^{(2)}\ldots $$ In general, considering the functor, $\mathfrak{M}(\quad ,n)$, from the category of simplicial groups to that of crossed $n$-cubes, gives the corresponding morphisms $$ \mathfrak{M}({\bf{F}}^{(0)},\ n) \rightarrow \mathfrak{M}({\bf \bf{F}}^{(1)},\ n) \rightarrow \mathfrak{M}({\bf{F}}^{(2)},\ n)\rightarrow ... \rightarrow \mathfrak{M}({\bf{F}},\ n). $$ We will investigate ${\bf \mathfrak{M}(\bf{F}}^{(i)}{\bf ,\ }n{\bf )}$, for $n=0,1,2, $ and varying $i$. Firstly look at $\mathfrak{M}({\bf{F}}^{(0)}{\bf ,\ }n{\bf ),\ }$ where the 0-skeleton ${\bf{F}}^{(0)}\,$ can be thought of as simplifying to $$\begin{array}{lccc} {\bf{F}}^{(0)}: & \cdots \longrightarrow {F({\bf B_1})}\longrightarrow {F({\bf B_1})}\longrightarrow {F({\bf B_1})} \end{array} $$ with the $d_i^n=s_j^n=\ $identity homomorphism on ${F({\bf B_1})}$. For $n=0,\,$ there is an equality $$ \mathfrak{M}({\bf{F}}^{(0)}{\bf ,\ }0{\bf )=}F_0^{(0)}/d_1 (\text{Ker}d_0)= {F({\bf B_1})}, $$ and so $\mathfrak{M}({\bf{F}}^{(0)}{\bf ,\ }0)$ is just the free group of 0-simplices of $\bf{F}$. For $n=1$, $\mathfrak{M}({\bf{F}}^{(0)}{\bf ,}1{\bf )}$ is $ NF_1^{(0)}/\partial _2NF_2^{(0)}\longrightarrow F_0. $ It is easy to show that $NF_1^{(0)}/\partial _2NF_2^{(0)}$ is trivial and hence $$ \mathfrak{M}({\bf{F}}^{(0)}{\bf ,\ }1{\bf )\cong (}1\longrightarrow {F({\bf B_1})}). $$ For $n=2$, \ $\mathfrak{M}({\bf{F}}^{(0)}{\bf ,\ }2{\bf )}$ is the trivial crossed square $$ \diagram NF_2 / d^{3}_{3} (NF_3 )\ddto\rrto&&\mbox{\rm Ker }d^{1}_{0}\ddto &&1 \ddto\rrto&& 1\ddto \\ &&& = \\ \mbox{\rm Ker }d^{1}_{1}\rrto&&F_1&&1\rrto&&F({\bf B_1}). \enddiagram $$ Next look at $\mathfrak{M}({\bf{F}}^{(1)}{\bf ,\ }n{\bf )}$ and recall that the 1-skeleton{~} $\bf{F}^{(1)}$ is $$ \diagram {\bf{F}}^{(1)} :\hspace{.5cm} ...{~}{F(s_1s_0({\bf B_1})\cup s_0({\bf B_2})\cup s_1({\bf B_2}))} \rto<0.25ex> \rto<1ex> \rto<1.75ex>^{\hspace{2.3cm} d_0 ,d_1 ,d_2 } & {F(s_0({\bf B_1})\cup {\bf B_2})} \lto<0.75ex> \lto<1.50ex>^{\hspace{2.4cm} s_1,s_0} \rto<0.25ex>\rto<1ex>^{\qquad d_1, d_0} & {F({\bf B_1})} \lto<0.75ex>^{\qquad s_0}. \enddiagram $$ For $n=0$, $\mathfrak{M}({\bf{F}}^{(1)},\ 0)$ is $ F_0^{(1)}/d_1(\text{Ker}d_0)\cong {F({\bf B_1})}/{\partial_1 NF_1}, $ which is $\pi _0({\bf{F}}^{(1)})\cong \pi _0(\bf{F}).$ For $n=1$, we have that $$ \begin{array}{rcl} \mathfrak{M}({\bf\bf{F}}^{(1)}{\bf ,\ }1{\bf )} & = & (NF_1/\partial _2NF_2\longrightarrow F_0), \\ & = & \langle{\bf B_2}\rangle/P_1\longrightarrow {F({\bf B_1})}, \end{array} $$ which is a free crossed module. In fact this is the free crossed module on the (generalised) presentation $({\bf B_1};{\bf B_2}, f_2)$. As pointed out in \cite{bh}, it is often convenient to generalise the notion of a presentation $({\bf X}, {\bf R})$ with $${\bf R}\subset F(X)$$ to one with the map ${\bf R} \rightarrow F(X)$ specified and not necessarily monic. Thus if $f_2$ is injective, this is just a presentation $\cal P$ of $\pi_1({\bf\bf{F}})$. The kernel of this crossed module is then the module of identities of $\cal P$, again see \cite{bh}. For $n=2$, $NF_2^{(1)} =\langle s_1({\bf B_2})\rangle \cap\langle Z\rangle$, so $\mathfrak{M}({\bf{F}}^{(1)}{\bf ,\ }2{\bf )\ }$ simplifies to give (up to isomorphism), $$ \diagram NF_2 / d^{3}_{3} (NF_3 )\ddto\rto&\mbox{\rm Ker }d^{1}_{0}\ddto &&J\ddto\rto& \langle {\bf B_2}\rangle \ddto \\ && = \\ \mbox{\rm Ker }d^{1}_{1}\rto&G_1&& \langle Z\rangle \rto& F(s_0({\bf B_1})\cup {\bf B_2}) \enddiagram $$ which is a crossed square with $J = (\langle s_1({\bf B_2})\rangle \cap\langle Z\rangle )/P_2.$ Next look at $\mathfrak{M}({\bf{F}}^{(2)}{\bf ,\ }n{\bf ).\ }$ Recall the 2-skeleton ${\bf{F}}^{(2)}$ is $$ \diagram { \bf{F}}^{(2)}:\hspace{.5cm} ... {F(s_1s_0({\bf B_1})\cup s_0({\bf B_2})\cup s_1({\bf B_2})\cup{\bf B_3})} \rto<0.25ex> \rto<1ex> \rto<1.75ex>^{\qquad\hspace{1.8cm} d_0 ,d_1 ,d_2} & {F(s_0({\bf B_1})\cup {\bf B_2})} \lto<0.75ex> \lto<1.50ex>^{\qquad\hspace{1.8cm} s_1,s_0} \rto<0.25ex> \rto<1ex>^{\qquad d_1, d_0} & {F({\bf B_1})} \lto<0.75ex>^{\qquad s_0}. \enddiagram $$ The following can be easily obtained by direct calculation : for $n=0,$ $$ \mathfrak{M}({ \bf{F}}^{(2)},0) = F_0/d_1(\text{Ker}d_0)\cong \pi _0({\bf{F}}^{(2)})= \mathfrak{M}({\bf{F}}^{(1)},0); $$ for $n=1,$ $$ \mathfrak{M}( \bf{F}^{(2)} ,1) \cong \langle {\bf B_2}\rangle /P_1\longrightarrow {F({\bf B_1})}. $$ Finally, let $n=2.$ By an earlier result of this section, $\mathfrak{M}({\bf{F}}^{(2)}{\bf ,}2{\bf )\ }$ corresponds to the free crossed square, $$ \diagram NF_2 / d^{3}_{3} (NF_3 ) \ddto\rto &\mbox{\rm Ker }d^{1}_{0}\ddto &&J\ddto\rto & \langle {\bf B_2}\rangle \ddto \\ && = \\ \mbox{\rm Ker }d^{1}_{1}\rto&F_1&& \langle Z_2\rangle \rto & F(s_0({\bf B_1})\cup {\bf B_2}) \enddiagram $$ where $J$ is now $(\langle s_1({\bf B_2})\cup{\bf B_3}\rangle \cap\langle Z\cup{\bf B_3}\rangle )/P_2$ and $\langle Z_2\rangle $ is $\langle Z\cup{\bf B_3}\rangle$, so this reduces to the earlier case if ${\bf B_3}$ is empty. Thus we have the following relations $$ \mathfrak{M}({\bf{F}}^{(2)}{\bf ,\ }0{\bf )}=\mathfrak{M}({\bf{F}}^{(1)}{\bf ,\ } 0), \qquad \mathfrak{M}({\bf{F}}^{(2)}{\bf ,\ }1{\bf )}=\mathfrak{M}({\bf{F}}^{(1)} {\bf ,\ }1{\bf )} $$ but $ \mathfrak{M}({\bf{F}}^{(2)}, 2)$ and $\mathfrak{M}({\bf{F}}^{(3)}, 2) $ need not be the same due to the additional influence of ${\bf B_3}$. Of course it is clear that, in general: $$ \begin{array}{ccccc} {\mathfrak{M}}({\bf{F}}^{(i)},\ n) & = & {\mathfrak{M}}({\bf{F}}^{(i+1)},\ n) & \text{if} & i\geq n+1. \end{array} $$ \section{Squared Complexes} The authors and Z. Arvasi have defined $n$-crossed complexes in \cite{zat}. In this paper, we will only need the case $n= 2$, which had already been defined by Ellis in \cite{ellis2}. We shall follow him in calling these {\em squared complexes}. A squared complex consists of a diagram of group homomorphisms $$ \diagram & & & & N\drto^{\mu'}\\ \ldots \rto& C_4\rto^{\partial_4}&C_3\rto^{\partial_3}& L\urto^{\lambda'}\drto_{\lambda}&& P\\ & & & &M\urto_{\mu} \enddiagram $$ together with actions of $P$ on $L, N, M$ and $C_i$ for $i\geq 3,$ and a function $h : M\times N\longrightarrow L.$ The following axioms need to be satisfied.\\ (i) The square $\left(\spreaddiagramrows{-1.2pc} \spreaddiagramcolumns{-1.2pc} \def\ssize} \def\labelstyle{\ssize{\ssize} \def\labelstyle{\ssize} \diagram L \dto_{{\lambda'}} \rto^{\lambda} & N \dto^{\mu} \\ {M} \rto_{\mu'}& P \enddiagram\right)$ is a crossed square; \\ (ii) The group $C_n$ is abelian for $n \geq 3$; \\ (iii) The boundary homomorphisms satisfy $\partial_n\partial_{n+1} = 1$ for $n \geq 3,$ and $\partial_3(C_3)$ lies in the intersection $\mbox{ker}\lambda\cap \mbox{ker}{\lambda'};$\\ (iv) The action of $P$ on $C_n$ for $n \geq 3$ is such that ${\mu}{M}$ and ${\mu'}N$ act trivially. Thus each $C_n$ is a $\pi_0$-module with $\pi_0 = P/{\mu}{M}{\mu'}N$; \\ (v) The homomorphisms $\partial_n$ are $\pi_0$-module homomorphisms for $n \geq 3.$ This last condition does make sense since the axioms for crossed squares imply that $\mbox{ker}{\mu'}\cap \mbox{ker}{\mu}$ is a $\pi_0$-module. A morphism of squared complexes $$\Phi: (C_\ast,\left(\spreaddiagramrows{-1.2pc} \spreaddiagramcolumns{-1.2pc} \def\ssize} \def\labelstyle{\ssize{\ssize} \def\labelstyle{\ssize} \diagram L \dto_{{\lambda'}} \rto^{\lambda} & N \dto^{\mu} \\ {M} \rto_{\mu'}& P \enddiagram\right) )\longrightarrow ({C_\ast'},\left(\spreaddiagramrows{-1.2pc} \spreaddiagramcolumns{-1.2pc} \def\ssize} \def\labelstyle{\ssize{\ssize} \def\labelstyle{\ssize} \diagram {L'} \dto_{{\lambda'}} \rto^{\lambda} & {N'} \dto^{\mu} \\ {M'} \rto_{\mu'}& {P'} \enddiagram\right))$$ consists of a morphism of crossed squares $(\Phi_{L}, \Phi_{N}, \Phi_{M}, \Phi_{P})$, together with a family of equivariant homomorphisms $\Phi_n$ for $n \geq 3$ satisfying $\Phi_{L}\partial_3 = {\partial'}_3\Phi_{L}$ and $\Phi_{n-1}\partial_n = {\partial'}_n\Phi_n$ for $n \geq 4.$ There is clearly a category $\mathfrak{SqComp}$ of squared complexes. This exists in both group and groupoid based versions. By a {\em (totally) free squared complex}, we will mean one in which the crossed square is (totally) free, and in which each $C_n$ is free as a $\pi_0$-module for $i\geq 3.$ \begin{prop} There is a functor $${\mathcal C}(\quad ,2) : \mathfrak{SimpGrp}\longrightarrow \mathfrak{SqComp} $$ such that free simplicial groups are sent to totally free squared complexes. \end{prop} \begin{pf} Let ${\bf G}$ be a simplicial group or groupoid. We will define a squared complex ${\mathcal C}({\bf G},2)$ by specifying ${\mathcal C}({\bf G},2)_A$ for each $A\subseteq <2>$ and for $n \geq 3$, ${\mathcal C}({\bf G},2)_n$. As usual, (cf. the other papers in this series, \cite{mp, mp1, mp2,mp3}), we will denote by $D_n$ the subgroup or subgroupoid of $NG_n$ generated by the degenerate elements. For $A\subset <2>,$ we define $${\mathcal C}({\bf G},2)_A = \mathfrak{M}(\mathfrak{sk_{2}}{\bf G,}2)_A = \frac{\cap\{\mbox{Ker}d_{i}^{2} : i \in A \}}{d_3(\mbox{Ker}d^3_0 \cap \bigcap\{\mbox{Ker}d_{i+1}^{3} : i \in A \} \cap D_3)}. $$ We do not need to define $\mu_i$ and the $h$-maps relative to these groups as they are already defined in the crossed square $\mathfrak{M}(\mathfrak{sk_{2}}{\bf G,}2)$. For $n \geq 3$, we set $${\mathcal C}({\bf G},2)_n = \frac{NG_n}{(NG_n\cap D_n)d_{n+1}(NG_{n+1}\cap D_{n+1})}.$$ As this is part of the crossed complex associated to ${\bf G}$, we can take the structure maps to be those of that crossed complex, cf. \cite{ep,mp2}. The terms are all modules over the corresponding $\pi_0$ as is easily checked. The final missing piece, $\partial_3$, of the structure is induced by the differential $\partial_3$ of $NG$. The axioms for a squared complex can now be verified using the known results for crossed squares and for crossed complexes with a direct verification of those axioms relating to the interaction of the two parts of the structure, much as in \cite{ep} and \cite{mp2}. Now suppose the simplicial group is free. The proof above of the freeness of $\mathfrak{M}(\mathfrak{sk_{2}}{\bf G,}2)$ together with the freeness of the crossed complex of a free simplicial group, \cite{mp2}, now completes the proof. \end{pf} \medskip Suppose that $\rho$ is a general squared complex. The {\em homotopy~groups} $\pi_{n}(\rho),$ $n\geq 0$ of $\rho$ are defined cf. \cite{ellis2}, to be the homology groups of the complex $$ \diagram \ldots\rto^{\partial_5~}&C_4\rto^{\partial_4~~}&C_3\rto^{\partial_2~}&L\rto^{\partial_2~\quad} &M\rtimes N\rto^{\quad\partial_1}&P\rto&1 \enddiagram $$ with $\partial_2(l)=({\lambda'}l^{-1}, \lambda l)$ and $\partial_1(m,n) = \mu(m){\mu'}(n).$ The axioms of a crossed square guarantee that $\partial_2$ and $\partial_1$ are homomorphisms with $\partial_3(C_3)$ normal in $\mbox{Ker}(\partial_2),~ \partial_2(L)$ normal in $\mbox{Ker}(\partial_1),$ and $\partial_1(M\rtimes N)$ normal in $P$. \begin{prop} The homotopy groups of ${\mathcal C}({\bf G},2)$ are isomorphic to those of ${\bf G}$ itself. \end{prop} \begin{pf} Again this is a consequence of well-known results on the two parts of the structure. \end{pf} \section{Alternative Descriptions of Freeness.} In the context of CW-complexes, Ellis, \cite{ellis2} gave a neat description of the top group $L$ in a (totally) free crossed square derived from that data. A simplicial group with a given CW-basis is the algebraic analogue of a CW-complex so one would expect a similar result to hold in that setting. Ellis uses the generalised van Kampen theorem of Brown and Loday, \cite{bl1}. In the algebraic setting no such tool is available, but in fact its use is not needed. Ellis' description is in terms of tensor products and coproducts. For completeness we recall the background definitions of these constructions. \subsection{Tensor Products} Suppose that $\mu: M\to P$ and $\nu: N\to P$ are crossed modules over $P.$ The groups $M$ and $N$ act on each other, and themselves, via the action of $P.$ The tensor product $M\otimes N$ is the group generated by the symbols $m\otimes n$ for $m\in M$, $n\in N$ subject to the relations $$mm'\otimes n = ({}^{m}{m'}\otimes {}^{m}n)(m\otimes n),$$ $$m\otimes n{n'} = (m\otimes n)({}^{n}{m}\otimes {}^{n}{n'}),$$ for $m,{m'}\in M, \ n,{n'}\in N.$ There are homomorphisms $\lambda: M\otimes N \to M, \ {\lambda'}: M\otimes N \to N$ defined on generators by $\lambda(m\otimes n)= m({}^{n}m)^{-1}$ and ${\lambda'}(m\otimes n)= ({}^{m}n)n^{-1}.$ The group $P$ acts on $M\otimes N$ by ${}^{p}(m\otimes n) = ({}^{p}m\otimes {}^{p}n),$ and there is a function $h: M\times N \to M\otimes N,$ $(m,n)\longmapsto m\otimes n.$ In \cite{bl1}, it is verified that this structure gives a crossed square $$ \diagram M\otimes N \rto^{\lambda} \dto_{\lambda'} & N\dto^{\nu} \\ M \rto_{\mu} & P \\ \enddiagram $$ with the universal property of extending the corner $$ \diagram & N\dto^{\nu} \\ M \rto_{\mu} & P \\ \enddiagram. $$ \subsection{Coproducts} Let $(M, P,\partial_1), (N,P,\partial_2)$ be $P$-crossed modules. Then $N$ acts on $M,$ and $M$ acts on $N,$ via the given actions of $P.$ Let $M\rtimes N$ denote the semidirect product with the multiplication given by $$ (m,n)({m'},{n'})=(m{m'},{~} {}^{m'}n{n'}) $$ and injections $$ \begin{array}{cc} {i'}: M\to M\rtimes N \qquad \mbox{and} \qquad {j'} : N\to M\rtimes N \\ \quad m\longmapsto (m,1)\qquad{~}\qquad\qquad\qquad n\longmapsto(1,n). \end{array} $$ We define the pre-crossed module $$ \begin{array}{cc} \underline{\delta}: M\rtimes N\to P \\ (m,n) \longmapsto \partial_1(m)\partial_2(n). \end{array} $$ Let $\{M, N\}$ be the subgroup of $M\rtimes N$ generated by the elements of the form $$ (m{}^nm^{-1},n{}^mn^{-1}) $$ for all $m\in M$, $n\in N$, thus we are able to form the quotient group $M\rtimes N/\{M,N\}$ and obtain an induced morphism $$ \partial: M\rtimes N/\{M,N\}\to P $$ given by $$ \partial(m,n)\{M,N\} = \partial_1(m)\partial_2(n). $$ Let $q: M\rtimes N\to M\rtimes N/\{M,N\}$ be projection and let $i = q{i'}, \ j= q{j'}.$ Then $M\circ N = (M\rtimes N)/\{M,N\}$ with the morphisms $i,j,$ is {\em a coproduct} of $(M,P,\partial_1)$ and $(N,P,\partial_2)$ in the category of $P$-crossed modules. \begin{prop}~\cite{ellis2} Let $(L, M, \bar{M}, M\rtimes F)$ be a (totally) free crossed square on the $2$-dimensional construction data or on functions $(f_2, f_3)$ as described above. Let $\partial: C\to M\rtimes F$ be the free crossed module on the function ${\bf B_3}\to M\rtimes F$ given by $y\longmapsto (f_3y, 1).$ From the crossed module $M\otimes \bar{M} \to M\rtimes F$, then $L$ is isomorphic to the coproduct $(M\otimes \bar{M})\circ C$ factored by the relations $$ \begin{array}{cc} 1)\quad i(\partial c\otimes\bar{m}) = j(c)j({}^{\bar{m}}c^{-1}) \\ 2)\quad i(m\otimes\partial c) = j({}^{m}c)j(c^{-1}) \end{array} $$ for $c \in C, \ m\in M ~\mbox{and}~ \bar{m}\in\bar{M}.$ The homomorphisms $L\to M$, $ L\to \bar{M}$ are given by the homomorphisms $$ \lambda:M\otimes \bar{M}\to M\qquad\mbox{and}\qquad {\lambda'}: M\otimes \bar{M}\to \bar{M} $$ and $\partial : C\to M\cap\bar{M}.$ The $h$-map of the crossed square is given by $$ h(m, \bar{n}) = i(m\otimes\bar{n}) $$ for $m, n \in M.$ \end{prop} \begin{pf} This comes by direct verification using the universal properties of tensors and coproducts. \end{pf} \medskip \textbf{Remark: } For future applications it is again important to note that the result is not dependent on the crossed square being \emph{totally} free, although this is the form proved and used by Ellis, \cite{ellis2}. If $M\rightarrow F$ is any pre-crossed module, one can form the `corner' $$ \diagram & M\dto \\ \bar{M} \rto & M\rtimes F, \\ \enddiagram$$ complete it to a crossed square via $M\otimes \bar{M}$ and then add in ${\bf B_3}\to M$. Nowhere does this use freeness of $M\rightarrow F$. \begin{cor} Let $\bf{G}^{(1)}$ be the $1$-skeleton of a simplicial group. Then in the free crossed square $\mathfrak{M}(\bf{G}^{(1)}, 2)$ described above, $$ NG_2^{(1)}/\partial_3NG_3^{(1)}\cong \mbox{Ker}d_1^{1}\otimes\mbox{Ker}d_0^{1}. $$ \end{cor} \begin{pf} This is clear from the previous proposition. \end{pf} \medskip \textbf{Remark: } If we set $M = \mbox{Ker}d^1_0 = NG^{(1)}_1$, then the identification given by the Corollary gives $$NG^{(1)}_2/\partial_3NG^{(1)}_3 \cong M\otimes \bar{M}.$$ This uses the fact that $\mbox{Ker}d^1_0 $ and $\mbox{Ker}d^1_1$ are linked via the map sending $m$ to $ms_0d_1m^{-1}$ for $m \in \mbox{Ker}d^1_0 $. The $h$-map $h : M \times \bar{M} \rightarrow NG^{(1)}_2/d^3_3NG^{(1)}_3$ is $h(x,y) = [s_1x,s_1ys_0y^{- 1}]d^3_3NG^{(1)}_3$, but this is also $h(x,y) = x\otimes y$. Thus $$x\otimes y = [s_1x,s_1ys_0y^{-1}]d^3_3NG^{(1)}_3$$ under the identification via the isomorphism of 5.2. This explains the `mysterious' formula of \cite{mp} in the discussion before Proposition 4.6 of that paper. \subsection{Applications to 2-crossed complexes.} Of course there are similar results for free squared complexes. What is less obvious is the way in which these results can be applied to the situation that we studied in our earlier paper, \cite{mp3}. There we considered the alternative model for 3-types given by Conduch\'e's 2-crossed modules and also looked at the corresponding 2-crossed complexes. We will not repeat all that discussion here but note the definition: {\bf Definition:}\\ A 2-crossed complex of group(oid)s is a sequence of group(oid)s $$C:\hspace{1cm} \ldots \rightarrow C_n \stackrel{\partial_n}{\rightarrow}C_{n-1}\rightarrow\ldots C_2 \stackrel{\partial_2}{\rightarrow}C_1\stackrel{\partial_1}{\rightarrow}C_0$$ in which\\ (i) ~$C_n$ is abelian for $n\geq 3$;\\ (ii)~$C_0$ acts on $C_n$, $n\geq 1$, the action of $\partial C_1$ being trivial on $C_n$ for $n\geq 3$;\\ (iii)~ each $\partial_n$ is a $C_0$-group(oid) homomorphism and $\partial_i\partial_{i+1} =1$ for all $i\geq 1$;\\ and\\ (iv)~ $C_2 \stackrel{\partial_2}{\rightarrow}C_1\stackrel{\partial_1}{\rightarrow}C_0$ is a 2-crossed module. \medskip We refer the reader to \cite{conduche} or \cite{mp3} for the exact meaning of 2-crossed module. Given a simplicial group or groupoid, ${\bf G}$, define $$C_n = \left\{\begin{array}{ll} NG_n & \mbox{ \rm for } n = 0,1\\ NG_2/d_3(NG_3 \cap D_3)& \mbox{ \rm for } n = 2\\ NG_n/(NG_n \cap D_n)d_{n+1}(NG_{n+1} \cap D_{n+1})& \mbox{ \rm for } n \geq 3 \end{array}\right. $$ with $\partial_n$ induced by the differential of $\bf{NG}$. Note that the bottom three terms (for $n = $ 0, 1, and 2) form a 2-crossed module considered in \cite{conduche} or \cite{mp3} and that for $ n \geq 3$, the groups are all $ \pi_0(G)$-modules, since in these dimensions $C_n$ is the same as the corresponding crossed complex term (cf. Ehlers and Porter \cite{ep} for instance). \begin{prop}\cite{mp3} With the above structure $(C_n, \partial_n)$ is a 2-crossed complex, which will be denoted $C(\mathbf{G})$.\hfill$\Box$ \end{prop} Here we note in particular that the term $C_2$ is $ NG_2/d_3(NG_3 \cap D_3)$ and so is the same as ${\mathcal C}({\bf G},2)_{<2>}$. Thus if ${\bf G}$ is a simplicial group, we obtain {\em gratis}: \begin{cor} Let $\bf{G}^{(1)}$ be the $1$-skeleton of a simplicial group. The 2-crossed complex of $\bf{G}^{(1)}$ satisfies $$C(\mathbf{G}^{(1)})_2\cong \mbox{Ker}d_1^{1}\otimes\mbox{Ker}d_0^{1}. $$\hfill$\Box$ \end{cor} We also get in general a description of $C(\bf{G}^{(2)})_2$ as a quotient of the form $( \mbox{Ker}d_1^{1} \otimes\mbox{Ker}d_0^{1}\circ C)/\sim$ where as in Proposition 5.1, this $C$ is a free crossed module on the `new cells' in dimension 2. \subsection{The suspension of a $K(\pi,1)$.} As was mentioned in \cite{mp2}, Brown and Loday used their generalised van Kampen Theorem, \cite{bl1}, to calculate $\pi_3\Sigma K(\pi,1)$ for $\pi$ a group, as the kernel of the commutator map from $\pi \otimes \pi$ to $\pi$. Jie Wu, (\cite{wu} Theorem 5.9), for any group $\pi$ and set of generators $\{x_\alpha | \alpha \in J\}$ for $\pi$, gives a presentation of $\pi_n\Sigma K(\pi,1)$ in terms of higher commutators, but does not manage to get the Brown-Loday result explicitly although his result is clearly linked to theirs. Wu's methods use a study of simplicial groups and a construction he ascribes to Carlsson, \cite{carlsson}. This gives a simplicial group $F^\pi(S^1)$ that has $\pi_{n+2}\Sigma K(\pi,1)\cong \Omega \Sigma K(\pi,1) \cong \pi^{n+1} F^\pi(S^1)$. As we pointed out in \cite{mp}, $F^\pi(S^1)$ is a pointed analogue of the `tensorisation' of $K(\pi,0)$, the constant simplicial group on $\pi$, with the simplicial circle $S^1$. In general if $G$ is a simplicial group and $K$ a pointed simplicial set, $G\bar{\wedge}K$ will denote the simplicial group with group of $n$- simplices given by $$ \coprod\limits_{x\in K_n}(G_n)_{x}/(G_n)_{\ast}. $$ If $x \in K_n$, we denote the $x$-indexed copy of $g \in G_n$ within $(G\bar{\wedge}K)_n$ by $g\bar{\wedge} x$. The face and degeneracy maps of $G\bar{\wedge}K$ are induced by the componentwise application of the corresponding morphisms of $G$ and $K$ $$d_i(g\bar{\wedge}x) = d_i^Gg\bar{\wedge}d_i^Kx,$$ $$s_i(g\bar{\wedge}x) = s_i^Gg\bar{\wedge}s_i^Kx.$$ Of course if $d_i^Kx = \ast$ then $d_i(g\bar{\wedge}x) = 1$. The case of interest to us is $G = K(\pi,0)$, $K = S^1$ and we will adopt the notation for simplices in $S^1$ used by us in \cite{mp}. We write $S^1_0 = \{ \ast\}$ and will take $\ast$ to denote the corresponding degenerate $n$-simplex basing $S^1_n$ in all dimensions; $S^1_1 = \{\sigma, \ast\}$, $S^1_2 = \{x_0, x_1, \ast\}$, where $x_0 = s_1\sigma$, $x_1 = s_0\sigma$ and in general $S_{n+1}^1 = \{x_0, \ldots, x_n, \ast\}$, where $x_i = s_n\ldots s_{i+1}s_{i-1} \ldots s_0\sigma$, $0\leq i \leq n$. We write $G = K(\pi,0)$ for simplicity and will usually make no distinction between simplices in different dimensions unless confusion might arise. We have $(G\bar{\wedge}S^1)_0 = 1$, \quad the trivial group, $(G\bar{\wedge}S^1)_1 \cong \pi,$ $(G\bar{\wedge}S^1)_2 \cong \pi \ast \pi$, \qquad the free product of two copies of $\pi$, and so on. \\ The group $(G\bar{\wedge}S^1)_n$ is a free product of $n$-copies of $\pi$, $\coprod\{(\pi)_x : x \in S^1_n\setminus \{\ast\}\}$, and writing as above $g\bar{\wedge}x$ for the $x$-indexed copy of $g\in \pi$ in this, we note that $(g\bar{\wedge}x)(g^\prime\bar{\wedge}x) = (gg^\prime\bar{\wedge}x)$ for $g$, $g^\prime \in \pi$. As $g\bar{\wedge}x_i^{(n+1)} = s_n(g\bar{\wedge}x_i^{(n)})$ holds in all dimensions, $n\geq 2$ and for all $0 \leq i \leq n$, it is clear that $N(G\bar{\wedge}S^1)_n = D_n$, that is, it is generated by degenerate elements in all dimensions $n \geq 2$, we can therefore apply Corollary 5.2. As $N(G\bar{\wedge}S^1)_0$ is trivial, $\mbox{Ker}d_0^1 = \mbox{Ker}d_1^1 =(G\bar{\wedge}S^1)_1 \cong \pi$, so we get: For $H = G\bar{\wedge}S^1$,$$NH_2/\partial_3NH_3 \cong \pi \otimes \pi.$$ We have by \cite{porter} that the algebraic 2-type of $H$ is completely modelled by the crossed square $\mathfrak{M}(H,2)$, that is by $$ \diagram \pi\otimes\pi \rto^{\mu_2}\dto_{\mu_1}&\pi\dto^=\\ \pi\rto_=&\pi, \enddiagram $$ where $\mu_1$ and $\mu_2$ are the commutator maps. As a consequence we have: \begin{cor} The 3-type of $\Sigma K(\pi,1)$ is completely specified by the above crossed square. In particular there is an isomorphism $$\pi_3(\Sigma K(\pi,1)) \cong \mbox{Ker}(\mu : \pi \otimes \pi \rightarrow \pi).$$ \hfill$\Box$ \end{cor} This result was first found by Brown and Loday \cite{bl1}. Their proof was an illustration of the use of their generalised van Kampen Theorem. Jie Wu, \cite{wu}, gives some methods that shed light on the higher homotopy groups, but although they yield a description of $\pi_4$, they do not analyse the 4-type itself. The model $G\bar{\wedge}S^1$ is 1-skeletal and one might expect that a triple tensor $\pi \otimes\pi \otimes\pi$ may be involved in any model of its 4-type. Of course $\mathfrak{M}(H,3)$ gives a complete model, but the individual terms involved in that model are not as easy to analyse as in $\mathfrak{M}(H,2)$. An amalgam of Wu's methods and the methods developed in the earlier papers of this series, \cite{mp,mp1,mp2,mp3}, might provide insight into this. This problem is not of itself that important, but it does seem to provide an excellent testbed for the development of methods to aid in calculation with low dimensional algebraic models of homotopy types.
\section{Introduction} Neutrino oscillations~\cite{pon} play a central role in neutrino physics. The most important condition for this phenomenon is given by neutrino mixing which is described by $\nu_{L\alpha} = \sum_j U_{\alpha j} \nu_{Lj}$ with $\alpha = e, \mu, \tau, \ldots$ and $j = 1,2,3, \ldots$ labelling neutrino flavours (types) and mass eigenfields, respectively. All neutrino oscillation experiments are evaluated with the formula \cite{pon} \begin{equation} P_{\nu_\alpha\to\nu_\beta} (L/E_\nu) = \left| \sum_j U_{\beta j} U^*_{\alpha j} \exp \left( -i \frac{m^2_j L}{2E_\nu} \right) \right|^2 \,, \label{P} \end{equation} where $U$ denotes the unitary mixing matrix, $L$ the distance between source and detector and $E_\nu$ the neutrino energy. The neutrino masses $m_j$ are associated with the mass eigenfields $\nu_j$. It has been indicated in several publications that the standard derivation of Eq.~(\ref{P}) raises a number of conceptual questions (see, e.g., Ref.~\cite{rich} for a clear exposition). Some of these questions are solved by the wave packet approach~\cite{kayser} (see also the review~\cite{zralek} where a list of references can be found), however, the size and form of the wave packet is not determined in this approach and remains a subject to reasonable estimates. The idea has been put forward to include the neutrino production and detection processes into the consideration of neutrino oscillations.~\cite{rich} Such an approach can be realized with \emph{quantum mechanics} -- in which case the neutrinos with definite mass are unobserved intermediate states between the source and detection processes~\cite{rich} -- or with \emph{quantum field theory} where the massive neutrinos are represented by inner lines in a big Feynman diagram depicting the combined source -- detection process.~\cite{GS96,GSM99} In the following we will discuss the field-theoretical treatment. The aims and hopes of such an approach are the following: 1. The elimination of the arbitrariness associated with the wave packet approach, 2. the description of neutrino oscillations by means of the particles in neutrino production and detection which are really manipulated in an experiment, 3. a more complete and realistic description in order to find possible limitations of formula (\ref{P}) in specific experimental situations. Considering laboratory experiments, there are two typical situations for neutrino oscillation experiments. The first one is \emph{decay at rest} (DAR) of the neutrino source. Its corresponding Feyman diagram is depicted in Fig.~\ref{DAR}. The wave functions of the source and detector particles are localized (peaked) at $\vec{x}_S$ and $\vec{x}_D$, respectively. The other situation is \emph{decay in flight} (DIF) of the neutrino source as represented in Fig.~\ref{DIF} where it is assumed that a proton hits a target localized at $\vec{x}_T$. The detector particle sits again at $\vec{x}_D$ but the source is not localized. In both situations the distance between source and detection is given by $L = |\vec{x}_D-\vec{x}_S|$. Note that in the Feynman diagrams of Figs.~\ref{DAR} and \ref{DIF} the neutrinos with definite mass occur as inner lines. In the spirit of our approach, neutrino oscillation probabilities are proportional to the cross sections derived from the amplitudes represented by these diagrams. \section{Assumptions and the resulting amplitude} The further discussion is based on the following assumptions: \begin{enumerate} \renewcommand{\labelenumi}{\Roman{enumi}.} \item The wave function $\phi_D$ of the detector particle does not spread with time which amounts to \begin{equation} \phi(\vec{x},t) = \psi_D(\vec{x}-\vec{x}_D)\, e^{-i E_{DP}t} \,, \end{equation} where $E_{DP}$ is the sharp energy of the detector particle and $\psi_D(\vec{y})$ is peaked at $\vec{y}=\vec{0}$. \item The detector is sensitive to momenta (energies) and possibly to observables commuting with momenta (charges, spin). \item The usual prescription for the calcuation of the cross section is valid. \end{enumerate} With the amplitudes symbolized by Figs.~\ref{DAR} and \ref{DIF} the oscillation probabilities are obtained by \begin{equation}\label{Pav} \left\langle P_{\stackrel{\scriptscriptstyle (-)}{\nu}_{\hskip-3pt \alpha} \to \stackrel{\scriptscriptstyle (-)}{\nu}_{\hskip-3pt \beta}} \right\rangle_\mathcal{P} \propto \int dP_S \int_\mathcal{P} \frac{d^3 p'_{D1}}{2E'_{D1}} \cdots \frac{d^3 p'_{D{n_D}}}{2E'_{D{n_D}}} \: \left| \mathcal{A}_{\stackrel{\scriptscriptstyle (-)}{\nu}_{\hskip-3pt \alpha} \to \stackrel{\scriptscriptstyle (-)}{\nu}_{\hskip-3pt \beta}} \right|^ 2 \,. \end{equation} In this equation we have indicated the average over some region $\mathcal{P}$ in the phase space of the final particle of the detection process. If no final particle of the neutrino production process is measured then one has to integrate over the total phase space of these final states. By definition, at the source (detector) a neutrino $\stackrel{\scriptscriptstyle (-)}{\nu}_{\hskip-3pt \alpha}$ ($\stackrel{\scriptscriptstyle (-)}{\nu}_{\hskip-3pt \beta}$) is produced (detected) if there is a charged lepton $\alpha^{\pm}$ ($\beta^{\pm}$) among the final states. In perturbation theory with respect to weak interactions, according to the Feynman diagrams Figs.~\ref{DAR} and \ref{DIF} one has to perform integrations $\int d^4 x_1$, $\int d^4 x_2$ and $\int d^4 q$ corresponding to the Hamiltonian densities for neutrino production and detection and the propagators of the mass eigenfields, respectively. These integrations are non-trivial because $\psi_D$ and the source (target) wave functions are not plane waves, but are localized at $\vec{x}_D$ and $\vec{x}_S$ ($\vec{x}_T$), respectively. After having performed the integrations over $x_{1,2}$ and $q^0$, in the asymptotic limit $L \to \infty$ only the neutrinos on mass shell contribute to the amplitude~\cite{GS96} which can be written as~\cite{GS96,GSM99} \begin{equation}\label{ampinfty} \mathcal{A}^\infty_{\nu_\alpha\to\nu_\beta} = \sum_j \mathcal{A}^S_j \mathcal{A}^D_j U_{\beta j}U^*_{\alpha j} e^{i q_j L} \end{equation} with \begin{equation}\label{kin} E_D = \sum_{b=1}^{n_D} E'_{Db} - E_{DP} \quad \mbox{and} \quad q_j = \sqrt{E_D^2 - m^2_j} \,. \end{equation} $\mathcal{A}^S_j$ and $\mathcal{A}^D_j$ denote the amplitudes for production and detection, respectively, of a neutrino with mass $m_j$. Note that $E_D$ is the energy on the neutrino line in Figs.~\ref{DAR} and \ref{DIF} and it is independent of $m_j$. This is an immediate consequence of the assumptions in this section. Furthermore, due to the above-mentioned integrations and the asymptotic limit we obtain \begin{equation}\label{AD} \mathcal{A}^D_j \propto \widetilde{\psi}_D(-q_j \vec{\ell} + \vec{p}\,'_D) \quad \mbox{with} \quad \vec{\ell} = (\vec{x}_D-\vec{x}_S)/L \quad \mbox{and} \quad \vec{p}\,'_D = \sum_{b=1}^{n_D} \vec{p}\,'_{Db} \,, \end{equation} where $\widetilde{\psi}_D$ is the Fourier transform of $\psi_D$. \section{Results} The preceding discussion leads us to the conclusion that with the assumptions stated in Section 2 the neutrino mass eigenstates are characterized by the energy $E_\nu \equiv E_D$ and momenta $q_j$ (\ref{kin}). Thus they have all the same energy determined by the detection process, but the momenta are different. The summation over $E_D$ is incoherent, i.e., it occurs in the cross section (see Eq.~(\ref{Pav})), not in the amplitude (\ref{ampinfty}). In this sense there are no neutrino wave packets in experiments conforming with our assumptions. Note that it has been pointed out in~\cite{KNW96} that a coherent or incoherent neutrino energy spread cannot be distinguished in neutrino oscillation experiments. Since the neutrino energy can in principle be determined with arbitrary precision the coherence length can theoretically be increased solely by detector manipulations.~\cite{KNW96,GK98} From Eq.~(\ref{AD}) it follows that with $\Delta m^2 \equiv |m^2_j-m^2_k|$ the condition \begin{equation}\label{ACC} |q_j-q_k| \simeq \frac{\Delta m^ 2}{2E_D} \lesssim \sigma_D \quad \mbox{or} \quad \sigma_{xD} \lesssim \frac{1}{4\pi} L^\mathrm{osc} \end{equation} is necessary for neutrino oscillations, where $\sigma_D$ and $\sigma_{xD}$ are the widths of the wave function of the detector particle in momentum and coordinate space, respectively, and $L^\mathrm{osc}$ is the oscillation length.~\cite{kayser,rich,GS96} In realistic experiments condition (\ref{ACC}) holds because $\sigma_{xD}$ is a microscopic whereas $L^\mathrm{osc}$ a macroscopic quantity. For DAR an analoguous condition exists for the width of the neutrino source wave function. For more details of the field-theoretical approach to neutrino oscillations, for a consideration of the finite lifetime of the neutrino source in the case of DAR and for an application of the results to the LSND and KARMEN experiments we refer the reader to~\cite{GS96,GSM99}. We have shown that in these experiments effects of the finite liftetime can be neglected. For a discussion of DIF see~\cite{campagne}. Thus in the framework discussed here all corrections to Eq.~(\ref{P}) are negligible. Note that we have not taken into account or discussed the interaction of the final state particles in the source with the environment (``interruption of neutrino emission''), a possible intermediate range of the asymptotic limit $L \to \infty$ as found in~\cite{ioan} and the possibility that in some cases (e.g., the KARMEN experiment) it is not realistic to use the conventional procedure to calculate the cross section (\ref{Pav}) by taking the asymptotic limit of the final time to infinity.
\section{Introduction} A Forcing Axiom ${\bf FA}_\kappa({\mathbb P})$ for a forcing notion ${\mathbb P}$ is a statement guaranteeing existence of directed subsets of ${\mathbb P}$ that meet any member of a pregiven family of size $\kappa$ of dense subsets of ${\mathbb P}$. As an axiom, ${\bf FA}_\kappa({\mathbb P})$ is a powerful combinatorial tool that allows one to get some of the properties of forcing extensions. Even more interesting are axioms demanding ${\bf FA}_\kappa({\mathbb P})$ for all forcing notions in a fixed family ${\mathcal K}$. Here the most popular are Martin's Axiom {{\bf MA}} postulating ${\bf FA}_\kappa({\mathbb P})$ for each ccc partial order ${\mathbb P}$ and any cardinal $\kappa<\con$ and (capturing more forcing notions) Proper Forcing Axiom {\bf PFA} postulating ${\bf FA}_{\aleph_1}({\mathbb P})$ for all proper forcing notions. The quest for giving the largest possible class of forcing notions ${\mathbb P}$ for which the axiom ${\bf FA}_{\aleph_1}({\mathbb P})$ may (simultaneously) hold was successfully accomplished by Foreman, Magidor and Shelah \cite{FMSh:240}, \cite{FMSh:252}, who introduced Martin's Maximum. In the present paper we want to start investigations in the opposite directions, looking in some sense for the {\em minimal} version of the standard Martin's Axiom. That is, we would like to have a model in which $\neg {\bf CH} + {\bf MA}$ holds but ${\bf FA}_{\aleph_1}({\mathbb P})$ fails for as many (necessarily not-ccc) forcing notions as possible. Our attention concentrates on forcing notions built according to the scheme of {\em norms on possibilities} of Ros{\l}anowski and Shelah \cite{RoSh:470}. (Unfortunately, some familiarity with that paper has to be assumed. In particular, for all definitions related to norms on possibilities we have to refer the reader to \cite{RoSh:470}.) These lines of investigations have some history already. Steprans proved that $\neg {\bf CH} + {\bf MA}$ is consistent with the negation of the forcing axiom for the Silver forcing notion (see \cite[\S 2]{CRSW93}). Next, Judah, Miller and Shelah \cite{JMSh:372}, and Velickovic \cite{Vel91Pos}, showed that $\neg {\bf CH} + {\bf MA}$ does not imply the forcing axiom for the Sacks forcing notion. It has been a common believe that the arguments of \cite{JMSh:372} can be repeated for a number of forcing notions in which conditions are finitely branching trees. For example, Brendle wrote in the proof of \cite[Proposition 5.1(c)]{Br96a} (about the method of \cite{JMSh:372}): ``it is easy to see that this argument works for any forcing notion with compact trees which doesn't have splitting going on every level''. It seems that \cite{RoSh:470} has provided the right formalism for specifying which forcing notions can be taken care of in this context. However, one would like to cover all of $\baire$--bounding forcing notions from \cite{RoSh:470} (so avoid the limitations on splittings which in the language of \cite{RoSh:470} would restrict us to t--omittory trees) and get the failure of ${\bf FA}_{\aleph_1}$ for all these forcing notions simultaneously. As remarked by Brendle, the obvious modifications of the method of \cite{JMSh:372} seem to be not applicable here. Therefore, we rather follow the Steprans way slightly generalizing it to be able to deal with a number of forcing notions in the same model. \medskip \noindent{\bf Notation}\quad Our notation is rather standard and compatible with that of classical textbooks on Set Theory (like Bartoszynski and Judah \cite{BaJu95}). However in forcing we keep the older convention that {\em a stronger condition is the larger one}. \begin{notation} \label{notacja} \begin{enumerate} \item For two sequences $\eta,\nu$ we write $\nu\vartriangleleft\eta$ whenever $\nu$ is a proper initial segment of $\eta$, and $\nu\trianglelefteq\eta$ when either $\nu\vartriangleleft\eta$ or $\nu=\eta$. The length of a sequence $\eta$ is denoted by $\lh(\eta)$. \item The cardinality of the continuum is denoted by $\con$. \item For a forcing notion ${\mathbb P}$, $\Gamma_{\mathbb P}$ stands for the canonical ${\mathbb P}$--name for the generic filter in ${\mathbb P}$. With this one exception, all ${\mathbb P}$--names for objects in the extension via ${\mathbb P}$ will be denoted with a dot above (e.g.~$\dot{s}$, $\dot{f}$). \item Ordinal numbers are denoted by $\alpha,\beta,\gamma,\delta, \varepsilon, \xi,\zeta$ (with possible indexes); cardinals will be called $\kappa,\mu$. The first infinite ordinal is $\omega$. The letters $u,v,\eta, \nu,\rho$ will stand for finite sequences. \end{enumerate} \end{notation} \noindent{\bf Where are the respective definitions in \cite{RoSh:470}?}\quad As we stated before, we have to assume that the reader is familiar with \cite{RoSh:470}. (Otherwise we would have to give the list of all needed definitions and it could be longer than the rest of the paper.) However, for reader's convenience we list below exact pointers to the descriptions of the cases of norms on possibilities that are used here. \begin{itemize} \item Weak creatures and weak creating pairs: \cite[\S 1.1]{RoSh:470}, in particular \cite[Definitions 1.1.1, 1.1.3]{RoSh:470}. \item Creatures and creating pairs: \cite[\S 1.2]{RoSh:470}, in particular \cite[Definitions 1.2.1, 1.2.2, 1.2.4, 1.2.5]{RoSh:470} (as in \cite{RoSh:470}, we assume here that all creating pairs are nice and smooth). \item Forcing notions ${\mathbb Q}^*_f(K,\Sigma)$, ${\mathbb Q}^*_{{\rm w}\infty}$: \cite[Definitions 1.1.6, 1.1.7, 1.1.10]{RoSh:470}. \item Tree--creatures and tree--creating pairs: \cite[\S 1.3]{RoSh:470}, in particular \cite[Definitions 1.3.1, 1.3.3]{RoSh:470}. \item Forcing notions ${\mathbb Q}^{\rm tree}_0(K,\Sigma)$, ${\mathbb Q}^{\rm tree}_1(K,\Sigma)$: \cite[Def.\ 1.3.5]{RoSh:470}. \item Creating pairs which capture singletons: \cite[Def.\ 2.1.10, 1.2.5(3)]{RoSh:470}. \item $\bar{2}$--big (tree) creating pairs: \cite[Def.\ 2.2.1, 2.3.2]{RoSh:470}. \item Halving Property: \cite[Def.\ 2.2.7]{RoSh:470}. \item ${\bf H}$--fast function $f$: \cite[Def.\ 1.1.12]{RoSh:470}. \item Simple creating pairs, gluing creating pairs: \cite[Def.\ 2.1.7]{RoSh:470}. \item t--omittory tree--creating pairs: \cite[Def.\ 2.3.4]{RoSh:470}. \item Strongly finitary creating pairs / tree--creating pairs: \cite[Def.\ 1.1.3 + 3.3.4]{RoSh:470}. \end{itemize} \noindent{\bf The forcing notions we want to take care of.}\quad Here we specify the family of forcing notions for which we want to get the failure of ${\bf FA}_{\aleph_1}$ (with keeping ${\bf MA}$). \begin{definition} Let $(K,\Sigma)$ be a weak creating pair for ${\bf H}$. We say that it is {\em typical} if for each $t\in K$ such that $(\exists u\in{\bf basis}(t))(|{\rm pos}(u,t)| =1)$ we have $\nor[t]\leq 1$. \end{definition} The reason for the above definition is the following. \begin{proposition} \label{divide} Suppose that $(K,\Sigma)$ is a 2-big creating pair (tree--creating pair, respectively), $t\in K$ is such that $\nor[t]\geq 4$. Let $u\in{\bf basis}(t)$. Then there are creatures (tree--creatures, resp.) $s_0,s_1\in\Sigma(t)$ such that $\nor[s_0],\nor[s_1]\geq \nor[t]-2$ and ${\rm pos}(u,s_0)\cap{\rm pos}(u,s_1)= \emptyset$. \end{proposition} \begin{proof} Choose a set $A\subseteq {\rm pos}(u,t)$ such that \begin{itemize} \item there is $s^*\in\Sigma(t)$ with ${\rm pos}(u,s^*)\subseteq A$ and $\nor[s^*]\geq\nor[t]-1$, but \item for each $a\in A$ and $s\in\Sigma(t)$, if ${\rm pos}(u,s)\subseteq A \setminus\{a\}$ then $\nor[s]<\nor[t]-1$. \end{itemize} Clearly it is possible; necessarily $|A|\geq 2$ (remember that $(K,\Sigma)$ is typical). Fix $a\in A$. Applying bigness to $s^*$ we get $s_0\in \Sigma(s^*)\subseteq\Sigma(t)$ such that $\nor[s_0]\geq\nor[t]-2$ and ${\rm pos}(u,s_0)\subseteq A\setminus\{a\}$. On the other hand, by the choice of the set $A$ (and bigness) we find $s_1\in\Sigma(t)$ such that $\nor[s_1]\geq \nor[t]-1$ and ${\rm pos}(u,s_1)\subseteq {\rm pos}(u,t)\setminus (A\setminus\{a\})$. \end{proof} \begin{definition} \label{classK} Let ${\mathcal K}$ be the family of all non-trivial forcing notions of one of the following types: \begin{enumerate} \item ${\mathbb Q}^{\rm tree}_1(K,\Sigma)$, where $(K,\Sigma)\in {\mathcal H}(\aleph_1)$ is a strongly finitary 2-big typical tree--creating pair for a function ${\bf H}\in{\mathcal H}(\aleph_1)$; \item ${\mathbb Q}^{\rm tree}_1(K,\Sigma)$, where $(K,\Sigma)\in {\mathcal H}(\aleph_1)$ is a strongly finitary typical t-omittory tree--creating pair for a function ${\bf H}\in{\mathcal H}(\aleph_1)$ (note that in this case ${\mathbb Q}^{\rm tree}_1(K, \Sigma)$ is a dense subforcing of ${\mathbb Q}^{\rm tree}_0(K,\Sigma)$, see \cite[2.3.5]{RoSh:470}); \item ${\mathbb Q}^*_f(K,\Sigma)$, where $(K,\Sigma)\in {\mathcal H}(\aleph_1)$ is a strongly finitary typical creating pair for ${\bf H}\in {\mathcal H}(\aleph_1)$, $f:\omega\times\omega\longrightarrow\omega$ is an ${\bf H}$-fast function, $(K, \Sigma)$ is $\bar{2}$-big, has the Halving Property and is either simple or gluing; \item ${\mathbb Q}^*_{{\rm w}\infty}(K,\Sigma)$, where $(K,\Sigma)\in {\mathcal H}(\aleph_1)$ is a strongly finitary typical creating pair for ${\bf H}\in{\mathcal H}(\aleph_1)$, $(K,\Sigma)$ captures singletons. \end{enumerate} \end{definition} \begin{theorem} [See {\cite[\S 2.3]{RoSh:470}}] All forcing notions in the class ${\mathcal K}$ are proper and $\baire$--bounding. \end{theorem} Now we may state the main result of this paper. \begin{theorem} \label{main} Assume $\kappa=\kappa^{<\kappa}$. Then there is a ccc forcing notion ${\mathbb P}$ of size $\kappa$ such that in $\bV^{{\mathbb P}}$: \begin{itemize} \item $\con=\kappa$ and ${\bf MA}$ holds true, but \item ${\bf FA}_{\aleph_1}({\mathbb Q})$ fails for every forcing notion ${\mathbb Q}\in{\mathcal K}$. \end{itemize} \end{theorem} The class ${\mathcal K}$ includes all $\baire$--bounding forcing notions presented in \cite{RoSh:470} (modulo the demand that they are additionally required to be in ${\mathcal H}(\aleph_1)$). In particular, the following forcing notions are in ${\mathcal K}$: \begin{enumerate} \item Sacks-like forcings. Let $\{I_n: n \in \omega\}$ be a sequence of finite sets. Every perfect set $P \subseteq \prod_n I_n$ corresponds to a tree $T$. Consider the forcing notion ${\mathbf P}$ which consists of trees $T$ such that $$ \forall s \in T \ \exists t \supseteq s \ \suc(t)=I_{|t|},$$ ordered by inclusion. If $I_n=2$ for all $n$ we get the Sacks forcing $\mathbf S$. If $I_n=k$ ($n \in \omega $) then we get forcing notions ${\mathbb D}_k$ (for $k<\omega$) of Newelski and Ros{\l}anowski \cite{NeRo93}. Finally, if $I_n=h(n)$ then we get the Shelah forcing ${\mathbf Q}_h$ of \cite{GJS92}. \item Silver-like forcings. For $h \in \omega^\omega$ let ${\mathbf P}_h=\{f: \omega \setminus \dom(f)\in [\omega]^\omega, \ \forall n \ (n \in \dom(f) \rightarrow f(n)\leq h(n))\}$. For $f,g \in {\mathbf P}_h$, $f \geq g$ if $g \subseteq f$. If $h(n)=2$ ($ n \in \omega $) then ${\mathbf P}_h$ is Silver forcing and for $h(n)=n$ we get Miller forcing from \cite{Mil81Som}. \item forcing notion ${\mathbb Q}_{f,g}$ from \cite{BJSh:368} its siblings from \cite{CiSh:653} and \cite{GoShMan93}. Suppose that $f \in \omega^\omega$ and $g \in \omega^{\omega \times \omega}$ are two functions such that \begin{enumerate} \item $f(n) > \prod_{j<n} f(j)$ for $ n \in \omega$, \item $g(n,j+1) > f(n) \cdot g(n,j)$ for $n,j \in \omega$, and \item $\min\left\{j \in \omega: g(n,j) > f(n+1)\right\} \stackrel{n \rightarrow \infty}{\longrightarrow} \infty$. \end{enumerate} Define $Seq^{f} = \{s \in \omega^{<\omega}: \forall j < |s|\ s(j)\leq f(j)\}$ and let $T \in {\mathbf {PT}}_{f,g}$ if \begin{enumerate} \item $T$ is a perfect subtree of $Seq^{f}$, and \item there exists a function $r \in \omega^{\omega}$, $\lim_{n \rightarrow \infty} r(n)=\infty$ such that $$\forall s \in T \ (stem(T) \subseteq s \rightarrow |\suc_{T}(s)| \geq g(|s|,r(|s|))).$$ \end{enumerate} \item Forcing $\operatorname{{\mathbf {S}}}_{g, g^\star}$ from \cite{BaJu95}. Let $g, g^\star \in \omega^\omega$ be two strictly increasing functions such that $g(0)=0$, $g^\star(0)=1$ and $g(n) << g^\star(n) << g(n+1)$ for all $n>0$. For $n \in \omega$ let $$P_n = \left\{a \subseteq g(n+1): |a|=\frac{g(n+1)}{2^n}\right\}.$$ For a set $A \subseteq P_n$ define $$\nor(A)=\min\{|X|: \forall a \in A \ X \not \subseteq a\}.$$ Let $$\operatorname{{\mathbf {S}}}_{g, g^\star}=\left\{\<t_n: n \in \omega\>: \forall n \ t_n \in P_n\ \&\ \forall k\ \limsup_{n \rightarrow \infty} \frac{\nor(t_n)}{g^\star(n+1)^k}=\infty\right\}.$$ For $p=\<t^0_n: n \in \omega\>$ and $q=\<t^1_n: n \in \omega\>$ we define $p \geq q$ if $t^0_n \subseteq t^1_n$ for all $n$. \item More complicated forcing notions from \cite{FrSh:406} and \cite{BaSh607}. \end{enumerate} \section{${\mathcal S}$--families of good graphs} In this section we introduce a property of families of graphs that will be one of our main tools later. \begin{definition} \label{goodgraph} Let ${\mathcal U}$ be a countable basis of a topology on a set $X$, $A\subseteq X\times\omega_1$. A triple ${\mathcal G}=(A,{\mathcal U},E)$ is a {\em good graph} if the following conditions are satisfied: \begin{enumerate} \item[(a)] $E\subseteq [A]^{\textstyle 2}$, and \item[(b)] if $(x_0,\alpha_0), (x_1,\alpha_1)\in A$ are distinct then $x_0\neq x_1$, and \item[(c)] if $(x_0,\alpha_0), (x_1,\alpha_1)\in A$, $x_0\neq x_1$ and $\{(x_0,\alpha_0),(x_1,\alpha_1)\}\notin E$ then there are disjoint $U_0,U_1\in{\mathcal U}$ such that $x_0\in U_0$, $x_1\in U_1$ and \[(\forall (x_0',\alpha_0'), (x_1',\alpha_1')\in A)(x_0'\in U_0\ \&\ x_1'\in U_1\quad\Rightarrow\quad\{(x_0',\alpha_0'), (x_1',\alpha_1')\}\notin E).\] \end{enumerate} \end{definition} \begin{definition} \label{Sfamily} Suppose that ${\mathcal F}=\{{\mathcal G}_\zeta:\zeta<\xi\}$ is a family of good graphs, ${\mathcal G}_\zeta=(A_\zeta,U_\zeta,E_\zeta)$. \begin{enumerate} \item Let $0<m<\omega$. An $m$--selector for ${\mathcal F}$ is a set $S\subseteq (\bigcup\limits_{\zeta<\xi}A_\zeta)^{\textstyle m}$ such that for some (not necessarily distinct) $\zeta_0,\ldots,\zeta_{m-1}<\xi$ we have \begin{enumerate} \item[$(\alpha)$] if $\nu\in S$, $\ell<m$ then $\nu(\ell)\in A_{\zeta_\ell}$, \item[$(\beta)$] if $\nu,\rho\in S$ are distinct, then for some $\ell<m$ we have $\{\nu(\ell),\rho(\ell)\}\in E_{\zeta_\ell}$. \end{enumerate} \item ${\mathcal F}$ is called an {\em ${\mathcal S}$--$m$-family} if there is no uncountable $m$--selector for ${\mathcal F}$. \item ${\mathcal F}$ is an {\em ${\mathcal S}$--family} (of good graphs) if it is an ${\mathcal S}$--$m$-family for each $m<\omega$. \end{enumerate} \end{definition} Let us show how we are going to use ${\mathcal S}$--families of good graphs. \begin{definition} \label{represent} Let ${\mathbb P}=({\mathbb P},\leq)$ be a forcing notion and let ${\mathcal G}=(A,{\mathcal U},E)$ be a good graph (with $A\subseteq X\times\omega_1$, ${\mathcal U}$ a countable basis of a topology on $X$). We say that ${\mathcal G}$ is {\em densely representable by $\bot_{{\mathbb P}}$} if there is a one-to-one mapping $\pi:A\longrightarrow {\mathbb P}$ such that \begin{enumerate} \item[(a)] for each $\alpha<\omega_1$, the set $\{\pi(x,\alpha): (x,\alpha)\in A\}$ is dense in ${\mathbb P}$, and \item[(b)] if $(x_0,\alpha_0),(x_1,\alpha_1)\in A$ are distinct, $\{(x_0, \alpha_0),(x_1,\alpha_1)\}\notin E$ then the conditions $\pi(x_0,\alpha_0)$, $\pi(x_1,\alpha_1)$ are incompatible (in ${\mathbb P}$). \end{enumerate} \end{definition} \begin{proposition} \label{notFAP} Let ${\mathcal F}$ be an ${\mathcal S}$--$1$--family of good graphs. Suppose that ${\mathbb P}$ is a forcing notion such that some ${\mathcal G}\in{\mathcal F}$ is densely representable by $\bot_{{\mathbb P}}$. Then ${\bf FA}_{\aleph_1}({\mathbb P})$ fails. \end{proposition} \begin{proof} Let ${\mathcal G}=(A,{\mathcal U},E)$ and let $\pi:A\longrightarrow{\mathbb P}$ witness that ${\mathcal G}$ is densely representable by $\bot_{{\mathbb P}}$. Let $H_\xi=\{\pi(x,\xi):(x, \xi)\in A\}$ (for $\xi<\omega_1$). Then the sets $H_\xi$ are dense in ${\mathbb P}$ by \ref{represent}(a). We claim that they witness the failure of ${\bf FA}_{\aleph_1}({\mathbb P})$. So suppose that $G\subseteq{\mathbb P}$ is a directed set which meets each $H_\xi$. For every $\xi<\omega_1$ choose $(x_\xi,\xi)\in A$ such that $\pi(x_\xi,\xi)\in G\cap H_\xi$. Look at the set $S=\{\langle( x_\xi,\xi)\rangle:\xi<\omega_1\}$ --- it is an uncountable 1--selector from ${\mathcal F}$, a contradiction. \end{proof} One of problems that we will have to take care of when building the forcing notion needed for \ref{main} is preserving ``being an ${\mathcal S}$--family''. A part of this difficulty will be dealt with by ``killing the ccc of bad forcing notions''. \begin{proposition} \label{notccc} Let ${\mathcal F}$ be an ${\mathcal S}$--family of good graphs. Suppose that ${\mathbb P}$ is a ccc forcing notion such that \[\forces_{{\mathbb P}}\mbox{`` ${\mathcal F}$ is not an ${\mathcal S}$--family''.}\] Then there is a ccc forcing notion ${\mathbb P}^{{\mathcal F}}$ (of size $\aleph_1$) such that \[\forces_{{\mathbb P}^{{\mathcal F}}}\mbox{`` ${\mathbb P}$ is not ccc and ${\mathcal F}$ is an ${\mathcal S}$--family''.}\] \end{proposition} \begin{proof} Let ${\mathcal F}=\{{\mathcal G}_\zeta:\zeta<\xi\}$, ${\mathcal G}_\zeta=(A_\zeta,{\mathcal U}_\zeta,E_\zeta)$, ${\mathcal U}_\zeta$ a basis of a topology on $X_\zeta$, $A_\zeta\subseteq X_\zeta \times\omega_1$. Assume that some condition in ${\mathbb P}$ forces that ${\mathcal F}$ is not an ${\mathcal S}$--family. Then we find $m<\omega$, $\zeta_0,\ldots,\zeta_{m-1}< \xi$, $p\in{\mathbb P}$ and ${\mathbb P}$--names $\dot{\nu}_\alpha$ for elements of $(\bigcup\limits_{\ell<m}A_{\zeta_\ell})^{\textstyle m}$ (for $\alpha< \omega_1$) such that \[\begin{array}{ll} p\forces_{{\mathbb P}}&\mbox{`` } (\forall\alpha<\beta<\omega_1)(\dot{\nu}_\alpha \neq \dot{\nu}_\beta)\quad\&\quad (\forall\alpha<\omega_1)(\forall\ell<m) (\dot{\nu}_\alpha(\ell)\in A_{\zeta_\ell})\quad\&\\ \ &\ \ (\forall\alpha<\beta<\omega_1)(\exists\ell<m)(\{\dot{\nu}_\alpha( \ell),\dot{\nu}_\beta(\ell)\}\in E_{\zeta_\ell})\mbox{ ''.} \end{array}\] For each $\alpha<\omega_1$ choose a sequence $\nu_\alpha\in (\bigcup\limits_{\ell<m}A_{\zeta_\ell})^{\textstyle m}$ and a condition $p_\alpha\geq p$ such that $p_\alpha\forces\mbox{``} \dot{\nu}_\alpha= \nu_\alpha$''. (So necessarily $(\forall\alpha<\omega_1)(\forall\ell<m) (\nu(\ell)\in A_{\zeta_\ell})$.) Let ${\mathbb Q}$ be the following forcing notion: \smallskip {\bf a condition} in ${\mathbb Q}$ is a finite set $q\subseteq\omega_1$ such that \[(\forall\{\alpha,\beta\}\in [q]^{\textstyle 2})(\forall \ell<m)(\{\nu_\alpha(\ell),\nu_\beta(\ell)\}\notin E_{\zeta_\ell});\] {\bf the order} is the inclusion (i.e., $q_1\leq q_2$ if and only if $q_1\subseteq q_2$). \smallskip \noindent Note that if $\alpha,\beta\in q\in{\mathbb Q}$, $\alpha\neq\beta$ then the conditions $p_\alpha,p_\beta$ are incompatible (we will use it in \ref{Pnotccc}). \begin{claim} \label{trick} Assume $\{q_\varepsilon:\varepsilon<\omega_1\}\subseteq{\mathbb Q}$, $q_\varepsilon =\{\alpha^\varepsilon_i:i<k\}$ (the increasing enumeration; $k<\omega$). Then there is $Y\in [\omega_1]^{\textstyle \aleph_1}$ such that \begin{enumerate} \item[$(\otimes)_Y$] for each $\varepsilon,\delta\in Y$, if $q_\varepsilon, q_\delta$ are incompatible in ${\mathbb Q}$ then there are $i<k$ and $\ell<m$ such that $\{\nu_{\alpha^\delta_i}(\ell), \nu_{\alpha^\varepsilon_i}(\ell)\}\in E_{\zeta_\ell}$. \end{enumerate} \end{claim} \begin{proof}[Proof of the claim] For each $\varepsilon<\omega_1$, $\ell<m$ and distinct $i,j<k$ choose $U^{\varepsilon,i,j}_\ell\in {\mathcal U}_{\zeta_\ell}\cup\{*\}$ such that if $\nu_{\alpha^\varepsilon_i}(\ell)=\nu_{\alpha^\varepsilon_j}(\ell)$ then $U^{\varepsilon,i,j}_\ell= *$, otherwise $U^{\varepsilon,i,j}_\ell, U^{\varepsilon,j,i}\in {\mathcal U}_{\zeta_\ell}$ are such that \begin{enumerate} \item[$(\circledast_1)$] $\nu_{\alpha^\varepsilon_i}(\ell)\in U^{\varepsilon,i,j}_\ell\times\omega_1$,\quad $\nu_{\alpha^\varepsilon_j}( \ell)\in U^{\varepsilon,j,i}_\ell\times\omega_1$,\quad $U^{\varepsilon,i, j}_\ell\cap U^{\varepsilon,j,i}_\ell=\emptyset$, \quad and for all $(x_0, \alpha_0),(x_1,\alpha_1)\in A_{\zeta_\ell}$, \[x_0\in U^{\varepsilon,i,j}_\ell\ \&\ x_1\in U^{\varepsilon,j,i}_\ell\quad \Rightarrow\quad \{(x_0,\alpha_0),(x_1,\alpha_1)\}\notin E_{\zeta_\ell}.\] \end{enumerate} (Why possible? Remember the definition of ${\mathbb Q}$ and \ref{goodgraph}(c).) Each ${\mathcal U}_{\zeta_\ell}$ is countable, so there are $U^{i,j}_\ell\in {\mathcal U}_{\zeta_\ell}\cup\{*\}$ and an uncountable set $Y\subseteq\omega_1$ such that \begin{enumerate} \item[$(\circledast_2)$] $(\forall\varepsilon\in Y)(\forall i,j<k, i\neq j)( \forall\ell<m)(U^{\varepsilon,i,j}_\ell=U^{i,j}_\ell)$. \end{enumerate} Suppose that $\varepsilon,\delta\in Y$ and the conditions $q_\varepsilon, q_\delta$ are incompatible. It means that there are $i,j<k$ and $\ell<m$ such that $\{\nu_{\alpha^\varepsilon_i}(\ell),\nu_{\alpha^\delta_j}(\ell)\} \in E_{\zeta_\ell}$. We are going to show that we may demand $i=j$ (what will finish the proof of the claim). So suppose that $i\neq j$. If $\nu_{\alpha^\varepsilon_i}(\ell)\neq\nu_{\alpha^\varepsilon_j}(\ell)$, then (by $(\circledast_2)$) $\nu_{\alpha^\delta_i}(\ell)\neq\nu_{\alpha^\delta_j} (\ell)$ and (by $(\circledast_1)+(\circledast_2)$) $\nu_{\alpha^\delta_j}( \ell)\in U^{\delta,i,j}_\ell\times\omega_1=U^{\varepsilon,i,j}_\ell\times \omega_1$. But applying the last part of $(\circledast_1)$ we may conclude now that $\{\nu_{\alpha^\varepsilon_i}(\ell),\nu_{\alpha^\delta_j}(\ell)\} \notin E_{\zeta_\ell}$, a contradiction. So $\nu_{\alpha^\varepsilon_i}( \ell)=\nu_{\alpha^\varepsilon_j}(\ell)$. But then $\nu_{\alpha^\delta_i}( \ell)=\nu_{\alpha^\delta_j}(\ell)$ and thus $\{\nu_{\alpha^\varepsilon_i}( \ell),\nu_{\alpha^\delta_i}(\ell)\}\in E_{\zeta_\ell}$. \end{proof} \begin{claim} \label{Qccc} ${\mathbb Q}$ is a ccc forcing notion. \end{claim} \begin{proof}[Proof of the claim] Suppose that $\{q_\xi:\xi<\omega_1\} \subseteq{\mathbb Q}$ is an antichain in ${\mathbb Q}$. We may assume that, for some $k<\omega$, $|q_\xi|=k$ for all $\xi<\omega_1$. Let $q_\xi=\{\alpha^\xi_0, \alpha^\xi_1,\ldots,\alpha^\xi_{k-1}\}$ be the increasing enumeration. Using \ref{trick} we may find an uncountable $Y\subseteq\omega_1$ such that for each distinct $\varepsilon,\delta\in Y$ there are $i_{\varepsilon,\delta}<k$ and $\ell_{\varepsilon,\delta}<m$ with $\{\nu_{\alpha^\delta_{i_{\varepsilon, \delta}}}(\ell_{i_{\varepsilon,\delta}}),\nu_{\alpha^\varepsilon_{i_{ \varepsilon,\delta}}}(\ell_{i_{\varepsilon,\delta}})\}\in E_{\zeta_\ell}$. For each $\varepsilon\in Y$ let $\eta_\varepsilon$ be a sequence of length $k\cdot m$ such that \[\eta_\varepsilon(n)=\nu_{\alpha^\varepsilon_i}(\ell)\quad\mbox{ whenever } n=i\cdot m+\ell,\ i<k,\ \ell<m.\] Look at the set $\{\eta_\varepsilon:\varepsilon\in Y\}\subseteq (\bigcup\limits_{\zeta<\xi} A_\zeta)^{\textstyle k\cdot m}$. It should be clear that it is an uncountable $k\cdot m$--selector for ${\mathcal F}$ (the clause $(\beta)$ of \ref{Sfamily}(1) for $\varepsilon,\delta\in Y$ is witnessed by $i_{\varepsilon,\delta}\cdot m+ \ell_{\varepsilon,\delta}$). A contradiction. \end{proof} \begin{claim} \label{Qgood} $\forces_{{\mathbb Q}}$ `` ${\mathcal F}$ is an ${\mathcal S}$--family of good graphs ''. \end{claim} \begin{proof}[Proof of the claim] The only bad thing that could happen after forcing with ${\mathbb Q}$ is that an uncountable $m^*$--selector was added (for some $m^*<\omega$). If so, then we have $m^*<\omega$, ${\mathbb Q}$--names $\dot{\eta}_\varepsilon$ (for $\varepsilon<\omega_1$) and a condition $q\in{\mathbb Q}$ such that \[q\forces_{{\mathbb Q}}\mbox{`` }\{\dot{\eta}_\varepsilon:\varepsilon<\omega_1\} \mbox{ is an $m^*$--selector for ${\mathcal F}$,\ \ $\dot{\eta}_\varepsilon$'s are pairwise distinct ''.}\] Clearly we may require that for some $\zeta_0,\ldots,\zeta_{m^*-1}$, for each $\varepsilon<\omega_1$ the condition $q$ forces that $\dot{\eta}_\varepsilon (\ell)\in A_{\zeta_\ell}$ (for all $\ell<m^*$). For each $\varepsilon< \omega_1$ pick a sequence $\eta_\varepsilon$ and a condition $q_\varepsilon \geq q$ such that $q_\varepsilon\forces_{\mathbb Q}$``$\dot{\eta}_\varepsilon= \eta_\varepsilon$''. Next, choose an uncountable set $Y\subseteq\omega_1$ and $k<\omega$ such that for $\varepsilon\in Y$, $q_\varepsilon= \{\alpha^\epsilon_i:i<k\}$ (the increasing enumeration) and $(\otimes)_Y$ of \ref{trick} holds (possible by \ref{trick}). Let $\rho_\varepsilon$ (for $\varepsilon\in Y$) be sequences of length $k\cdot m+m^*$ defined by \[\rho_\varepsilon(n)=\left\{ \begin{array}{ll} \nu_{\alpha^\varepsilon_i}(\ell)&\mbox{if }n=i\cdot m+\ell,\ i<k,\ \ell<m,\\ \eta_\varepsilon(\ell) &\mbox{if }n=k\cdot m+\ell,\ \ell<m^*. \end{array}\right.\] Note that for distinct $\varepsilon,\delta\in Y$ we have: \begin{enumerate} \item[$(\boxtimes_1)$] if $q_\varepsilon,q_\delta$ are compatible in ${\mathbb Q}$, $\varepsilon<\delta<\omega_1$, then for some $\ell<m^*$ we have $\{\rho_\varepsilon(km+\ell),\rho_\delta(km+\ell)\}=\{\eta_\varepsilon(\ell), \eta_\delta(\ell)\}\in E_{\zeta_\ell}$; \item[$(\boxtimes)_2$] if $q_\varepsilon,q_\delta$ are incompatible in ${\mathbb Q}$ then there are $i<k$ and $\ell<m$ so that $\{\rho_\varepsilon(im+\ell), \rho_\delta(im+\ell)\}=\{\nu_{\alpha^\varepsilon_i}(\ell),\nu_{ \alpha^\delta_i}(\ell)\}\in E_{\zeta_\ell}$. \end{enumerate} (Why? $(\boxtimes_1)$ follows from the choice of $\dot{\eta}_\varepsilon$, $q_\varepsilon$, $(\boxtimes_2)$ is a consequence of $(\otimes)_Y$.) Hence, $\{\rho_\varepsilon:\varepsilon\in Y\}$ is an uncountable $km+m^*$--selector from ${\mathcal F}$, a contradiction. \end{proof} \begin{claim} \label{Pnotccc} For some $q\in{\mathbb Q}$ we have \[q\forces_{{\mathbb Q}}\mbox{ `` ${\mathbb P}$ does not satisfy the ccc ''.}\] \end{claim} \begin{proof}[Proof of the claim] As we stated before, if $q\in{\mathbb Q}$ and $\alpha,\beta\in q$ are distinct then the conditions $p_\alpha,p_\beta$ are incompatible in ${\mathbb P}$. Since, by \ref{Qccc}, the forcing notion ${\mathbb Q}$ is ccc, there is a condition $q\in{\mathbb Q}$ such that \[q\forces_{{\mathbb Q}}\mbox{`` }\{\alpha<\omega_1: \{\alpha\}\in\Gamma_{\mathbb Q}\}\mbox{ is uncountable ''}.\] It should be clear now that the condition $q$ forces (in ${\mathbb Q}$) that ${\mathbb P}$ fails the ccc. \end{proof} Let $q\in{\mathbb Q}$ be as guaranteed by \ref{Pnotccc} and let ${\mathbb P}^{\mathcal F}$ be the ${\mathbb Q}$ restricted to elements stronger than that $q$. It should be clear that ${\mathbb P}^{\mathcal F}$ is as required in the proposition. \end{proof} \begin{conclusion} \label{conclusion} If ${\mathcal F}$ is an ${\mathcal S}$--family of good graphs and ${\mathbb P}$ is a forcing notion with the Knaster property, then \[\forces_{{\mathbb P}}\mbox{`` ${\mathcal F}$ is an ${\mathcal S}$--family ''.}\] \end{conclusion} \section{Where are our ${\mathcal S}$--families from?} It follows from \ref{notFAP} that to make sure that ${\bf FA}_{\aleph_1}({\mathbb P})$ fails for all ${\mathbb P}\in {\mathcal K}$ it is enough to have an ${\mathcal S}$--family of good graphs such that for every ${\mathbb P}\in{\mathcal K}$ some ${\mathcal G}\in{\mathcal F}$ is densely representable by $\bot_{{\mathbb P}}$. In this section we will (almost) show how the respective ${\mathcal S}$--family is created in our model. Basically, it will come from Cohen reals, but interpreted in a special way. Let ${\mathbb P}\in{\mathcal K}$ and let $p$ be a condition in ${\mathbb P}$. Considering all possible cases (of \ref{classK}) we define a countable basis ${\mathcal U}({\mathbb P})$ of a topology on $X({\mathbb P})\stackrel{\rm def}{=}{\mathbb P}$, a set $E({\mathbb P})\subseteq [X({\mathbb P})\times\omega_1]^{\textstyle 2}$ and a forcing notion ${\mathbb C}({\mathbb P},p)$. (In the last case we will assume additionally that the condition $p$ has some special form, which will restrict us to a dense subset of ${\mathbb P}$.) \medskip \noindent{\sc Case 1}:\quad ${\mathbb P}={\mathbb Q}^{\rm tree}_1(K,\Sigma)$, where $(K,\Sigma)\in {\mathcal H}(\aleph_1)$ is a strongly finitary 2-big typical tree--creating pair for a function ${\bf H}\in{\mathcal H}(\aleph_1)$. \noindent For $q=\langle t^q_\eta:\eta\in T^q\rangle\in{\mathbb P}$ and $N<\omega$ let $U(q,N)$ be the family of all $q'\in{\mathbb P}$ such that \[\mrot(q')=\mrot(q)\quad\mbox{ and }\quad (\forall\eta\in T^{q'})\big(\lh (\eta)<N\quad\Rightarrow\quad (\eta\in T^q\ \&\ t^q_\eta=t^{q'}_\eta) \big).\] Let ${\mathcal U}({\mathbb P})$ consist of all nonempty sets $U(q,N)$ (for $q\in {\mathbb P}$, $N<\omega$). Clearly ${\mathcal U}({\mathbb P})$ is a countable basis of a topology on ${\mathbb P}$ (which is the natural product topology). Next we define \[\begin{array}{ll} E({\mathbb P})=\{\{(q_0,\alpha_0),(q_1,\alpha_1)\}\in [{\mathbb P}\!\times\!\omega_1]^{ \textstyle 2}:& \mbox{for each }N<\omega\\ \ &{\rm dcl}(T^{q_0})\cap \prod\limits_{i<N}{\bf H}(i)\cap{\rm dcl}(T^{q_1})\neq\emptyset\} \end{array}\] (where ${\rm dcl}(T)$ is the downward closure of a quasi tree $T$; see \cite[Def.\ 1.3.1]{RoSh:470}). The forcing notion ${\mathbb C}({\mathbb P},p)$ is defined as follows: \smallskip \noindent{\bf a condition $r$ in ${\mathbb C}({\mathbb P},p)$} is a finite system $r= \langle s^r_\eta: \eta\in \hat{S^r}\rangle$ such that $S^r\subseteq T^p$ is a (finite) quasi tree, $\mrot(S^r)=\mrot(T^p)$, \[(\forall\eta\in \hat{S^r})(s^r_\eta\in\Sigma(t^p_\eta)\ \&\ \nor[s^r_\eta]\geq\nor[t^p_\eta]-2\ \&\ {\rm pos}(s^r_\eta)=\suc_{S^r}(\eta)),\] and if $\nor[t^p_\eta]\leq 4$, $\eta\in \hat{S^r}$ then $s^r_\eta=t^p_\eta$; \noindent{\bf the order of ${\mathbb C}({\mathbb P},p)$} is the end extension, i.e., $r_0\leq r_1$ if and only if $S^{r_0}\subseteq S^{r_1}$ and $(\forall\eta\in \hat{S^{r_0}})(s^{r_0}_\eta=s^{r_1}_\eta)$. \smallskip \noindent It should be clear that ${\mathbb C}({\mathbb P},p)$ is a countable atomless (remember \ref{divide}) forcing notion, so it is equivalent to the Cohen forcing. Moreover, the forcing with ${\mathbb C}({\mathbb P},p)$ adds a condition $\dot{p}^*=\bigcup\Gamma_{{\mathbb C}({\mathbb P},p)}\in{\mathbb P}$ stronger than $p$. \medskip \noindent{\sc Case 2}:\quad ${\mathbb P}={\mathbb Q}^{\rm tree}_1(K,\Sigma)$, where $(K,\Sigma) \in {\mathcal H}(\aleph_1)$ is a strongly finitary typical t-omittory tree--creating pair for a function ${\bf H}\in{\mathcal H}(\aleph_1)$. \noindent ${\mathcal U}({\mathbb P})$ and $E({\mathbb P})$ are defined like in the Case 1. The forcing notion ${\mathbb C}({\mathbb P},p)$ is defined similarly too, though we now make an advantage from ``t--omittory'': \smallskip \noindent{\bf a condition $r$ in ${\mathbb C}({\mathbb P},p)$} is a finite system $r= \langle s^r_\eta: \eta\in \hat{S^r}\rangle$ such that $S^r\subseteq T^p$ is a quasi tree, $\mrot(S^r)=\mrot(T^p)$, and for each $\eta\in \hat{S^r}$ \begin{itemize} \item there is a (finite) quasi tree $T^*_\eta\subseteq T^p$ such that $\mrot(T^*_\eta)= \eta$ and $s^r_\eta\in\Sigma(t^p_\nu:\nu\in\hat{T^*_\eta })$, \item $\nor[s^r_\eta]\geq\min\{\nor[t^p_\nu]-2:\nu\in T^*_\eta\}$, \end{itemize} and if $\nor[t^p_\eta]\leq 4$, $\eta\in \hat{S^r}$ then $s^r_\eta=t^p_\eta$; \noindent{\bf the order of ${\mathbb C}({\mathbb P},p)$} is the end extension, so $r_0\leq r_1$ if and only if $S^{r_0}\subseteq S^{r_1}$ and $(\forall\eta\in \hat{S^{r_0}})(s^{r_0}_\eta=s^{r_1}_\eta)$. \smallskip \noindent Again, ${\mathbb C}({\mathbb P},p)$ is a countable atomless forcing notion. The forcing with ${\mathbb C}({\mathbb P},p)$ adds a condition $\dot{p}^*=\bigcup\Gamma_{{\mathbb C}( {\mathbb P},p)}\in{\mathbb P}$ stronger than $p$. \medskip \noindent{\sc Case 3}:\quad ${\mathbb P}={\mathbb Q}^*_f(K,\Sigma)$, where $(K,\Sigma)\in {\mathcal H}(\aleph_1)$ is a strongly finitary typical creating pair for ${\bf H}\in {\mathcal H}(\aleph_1)$, $f:\omega\times\omega\longrightarrow\omega$ is an ${\bf H}$-fast function, $(K,\Sigma)$ is $\bar{2}$-big, has the Halving Property and is either simple or gluing. \noindent For a condition $q=(w^q,t^q_0,t^q_1,\ldots)\in{\mathbb P}$ and $N<\omega$ we let $U(q,N)$ be the collection of all conditions $q'\in{\mathbb P}$ such that $w^q=w^{q'}$ and $t^q_n=t^{q'}_n$ for all $n<N$. Next, ${\mathcal U}({\mathbb P})$ is the family of all non-empty sets $U(q,N)$ (for $q\in{\mathbb P}$ and $N<\omega$). Like before, ${\mathcal U}({\mathbb P})$ is a countable basis of the standard (product) topology on ${\mathbb P}$. We define \[\begin{array}{ll} E({\mathbb P})&=\{\{(q_0,\alpha_0),(q_1,\alpha_1)\}\in [{\mathbb P}\times\omega_1]^{ \textstyle 2}: \mbox{ for each } N<\omega\\ \ &\ \ {\rm dcl}({\rm pos}(w^{q_0},t^{q_0}_0,\ldots,t^{q_0}_N))\cap \prod\limits_{i<N} {\bf H}(i)\cap{\rm dcl}({\rm pos}(w^{q_1},t^{q_1}_0,\ldots,t^{q_1}_N))\neq\emptyset\}. \end{array} \] The forcing notion ${\mathbb C}({\mathbb P},p)$ is defined so that \smallskip \noindent{\bf a condition $r$ in ${\mathbb C}({\mathbb P},p)$} is a finite sequence $r=(w^r, s^r_0,\ldots,s^r_{n_r})\in{\rm FC}(K,\Sigma)$ such that $w^r=w^p$, $s^r_k\in \Sigma(t^p_k)$, and \begin{itemize} \item if $\nor[t^p_k]>f(n,m^{t^p_k}_{\rm dn})$, $n\geq 4$ then $\nor[s^r_k]> f(n-2,m^{s^r_k}_{\rm dn})$, \item if $\nor[t^p_k]\leq f(4,m^{t^p_k}_{\rm dn})$ then $s^r_k=t^p_k$; \end{itemize} \noindent{\bf the order of ${\mathbb C}({\mathbb P},p)$} is the extension, i.e., $r_0\leq r_1$ if and only if $n_{r_0}\leq n_{r_1}$ and $s^{r_0}_k=s^{r_1}_k$ for all $k\leq n_{r_0}$. \smallskip \noindent Clearly ${\mathbb C}({\mathbb P},p)$ is countable and atomless (remember \ref{divide}). It adds a condition $\dot{p}^*=\bigcup\Gamma_{{\mathbb C}({\mathbb P},p)}\in {\mathbb P}$ stronger than $p$. \medskip \noindent{\sc Case 4}:\quad ${\mathbb P}={\mathbb Q}^*_{{\rm w}\infty}(K,\Sigma)$, where $(K,\Sigma)\in {\mathcal H}(\aleph_1)$ is a strongly finitary typical creating pair for ${\bf H}\in{\mathcal H} (\aleph_1)$, $(K,\Sigma)$ captures singletons. \noindent Both ${\mathcal U}({\mathbb P})$ and $E({\mathbb P})$ are defined like in Case 3. As we said before, here we will require that $p$ is of special form. Namely, we demand that $p=(w^p,t^p_0,t^p_2,\ldots)\in{\mathbb P}$ is such that for some strictly increasing sequence $\langle m_k:k<\omega\rangle\subseteq\omega$, $m_0=0$ and for each $k\in\omega$: \begin{itemize} \item $\nor[t^p_{m_k}]\geq 4+k$, and \item if $m_k+1<m_{k+1}$ then for some (equivalently: all) $u\in{\rm pos}(w^p, t^p_0,\ldots,t^p_{m_k})$ we have $|{\rm pos}(u,t^p_{m_k+1},\ldots, t^p_{m_{k+1} -1})|=1$. \end{itemize} (Since $(K,\Sigma)$ captures singletons the conditions of this form are dense in ${\mathbb P}$.) Now, the forcing notion ${\mathbb C}({\mathbb P},p)$ is defined so that \smallskip \noindent{\bf a condition $r$ in ${\mathbb C}({\mathbb P},p)$} is a finite sequence $r=(w^r, s^r_0,\ldots,s^r_{n_r})\in{\rm FC}(K,\Sigma)$ such that $w^r=w^p$, $s^r_\ell\in \Sigma(t^p_\ell)$, and \[(\forall k<\omega)(m_{2k}\leq n_r\ \Rightarrow\ \nor[s^r_{m_{2k}}]\geq \nor[t^p_{m_{2k}}]-2;\] \noindent{\bf the order of ${\mathbb C}({\mathbb P},p)$} is the extension, i.e., $r_0\leq r_1$ if and only if $n_{r_0}\leq n_{r_1}$ and $s^{r_0}_k=s^{r_1}_k$ for all $k\leq n_{r_0}$. \smallskip \noindent Again, ${\mathbb C}({\mathbb P},p)$ is countable and atomless, and it adds a condition $\dot{p}^*=\bigcup\Gamma_{{\mathbb C}({\mathbb P},p)}\in{\mathbb P}$ stronger than $p$. \begin{lemma} \label{oneforce} Suppose ${\mathbb P}\in{\mathcal K}$ and $p\in{\mathbb P}$ is such that ${\mathbb C}({\mathbb P},p)$ is defined. Let $r\in{\mathbb C}({\mathbb P},p)$. Then there are two conditions $r_0,r_1\in{\mathbb C}({\mathbb P},p)$ stronger than $r$ and basic open sets $U_0,U_1,U\in {\mathcal U}({\mathbb P})$ such that \begin{itemize} \item $U_0\cap U_1=\emptyset$, $p\in U$, \item if $p'\in U$ and ${\mathbb C}({\mathbb P},p')$ is defined then $r,r_0,r_1\in{\mathbb C}({\mathbb P}, p')$, $r_0,r_1$ are stronger than $r$ (in ${\mathbb C}({\mathbb P},p')$) and $r_0\forces_{{\mathbb C}({\mathbb P},p')}\dot{p}^*\in U_0$, $r_1\forces_{{\mathbb C}({\mathbb P},p')} \dot{p}^*\in U_1$, \item $(\forall q_0\in U_0)(\forall q_1\in U_1)(\forall\alpha_0,\alpha_1< \omega_1)(\{(q_0,\alpha_0),(q_1,\alpha_1)\}\notin E({\mathbb P}))$. \end{itemize} (Note that these formulas are absolute.) \end{lemma} \begin{proof} In Cases 1 and 3 (of \ref{classK}) use \ref{divide}; in other cases use directly the assumption that $(K,\Sigma)$ is typical. \end{proof} The ${\mathcal S}$--families in our model will be created by choosing sets $A\subseteq X({\mathbb P})\times \omega_1$ (for each ${\mathbb P}\in{\mathcal K}$) so that in each pair $(q,\alpha)\in A$ the first coordinate $q$ is added generically by the forcing notion ${\mathbb C}({\mathbb P},p)$ (for some condition $p\in{\mathbb P}$). For this we will use the finite support product of $\aleph_1$ copies of ${\mathbb C}({\mathbb P},p)$, denoted by ${\mathbb C}^{\omega_1}({\mathbb P},p)=\prod\limits_{\delta<\omega_1}{\mathbb C}({\mathbb P},p)$ (so a condition in ${\mathbb C}^{\omega_1}({\mathbb P},p)$ is a finite function $c:\dom(c) \longrightarrow{\mathbb C}({\mathbb P},p)$ and the order is the natural one). The forcing with ${\mathbb C}^{\omega_1}({\mathbb P},p)$ adds the set \[\dot{Z}^{\mathbb P}_p=\big\{(q,\alpha):\alpha<\omega_1\ \&\ q=\bigcup\{r: (\alpha,r)\in\Gamma_{{\mathbb C}^{\omega_1}({\mathbb P},p)}\}\big\}.\] (Sets of these form will be used to build a good graph ${\mathcal G}$ densely representable by $\bot_{{\mathbb P}}$.) \begin{definition} \label{isomorph} Suppose that ${\mathbb P}\in{\mathcal K}$, $p\in{\mathbb P}$ and ${\mathbb C}({\mathbb P},p)$ is defined. We say that two conditions $\bar{c}_0,\bar{c}_1\in{\mathbb C}^{\omega_1}({\mathbb P},p)$ are {\em isomorphic\/} (and then we write $\bar{c}_0\sim \bar{c}_1$) if $|\dom( \bar{c}_0)|=|\dom(\bar{c}_1)|$ and if $H:\dom(\bar{c}_0)\longrightarrow \dom(\bar{c}_1)$ is the order preserving bijection then $\bar{c}_1(H(\alpha ))=\bar{c}_0(\alpha)$ for each $\alpha\in\dom(\bar{c}_0)$. \end{definition} \noindent (Note that there are countably many isomorphism types of conditions in ${\mathbb C}^{\omega_1}({\mathbb P},p)$.) The main technical advantage of using the forcing notions ${\mathbb C}^{\omega_1} ({\mathbb P},p)$ to create our ${\mathcal S}$--families is presented by the following lemma. \begin{lemma} \label{maintech} Let ${\mathbb P}^0,\ldots,{\mathbb P}^k\in{\mathcal K}$. Suppose that $\langle {\mathbb P}_\xi,\dot{{\mathbb Q}}_\xi: \xi<\gamma\rangle$ is a finite support iteration of ccc forcing notions, $\gamma$ is a limit ordinal. Furthermore assume that for some disjoint sets $I_0,\ldots,I_k\subseteq\gamma$ we have \begin{enumerate} \item[$(\alpha)$] if $\xi\in I_\ell$ then $\forces_{{\mathbb P}_\xi}$`` $\dot{{\mathbb Q}}_\xi={\mathbb C}^{\omega_1}({\mathbb P}^\ell,p^\xi)$ for some $\dot{p}^\xi\in {\mathbb P}^\ell$'', \item[$(\beta)$] for $\zeta\leq\gamma$, $\dot{A}^\ell_\zeta$ is the ${\mathbb P}_\zeta$--name for the set $\bigcup\{\dot{Z}^{{\mathbb P}^\ell}_{\dot{p}^\xi}:\xi \in I_\ell\cap\zeta\}\subseteq{\mathbb P}^\ell\times \omega_1$, \item[$(\gamma)$] for each $\zeta<\gamma$ \[\forces_{{\mathbb P}_\zeta}\mbox{`` }\{(\dot{A}^\ell_\zeta, {\mathcal U}({\mathbb P}^\ell), E({\mathbb P}^\ell)\cap [\dot{A}^\ell_\zeta]^{\textstyle 2}): \ell\leq k\}\mbox{ is an ${\mathcal S}$--family of good graphs ''.}\] \end{enumerate} Then \[\forces_{{\mathbb P}_\gamma}\mbox{`` }\{(\dot{A}^\ell_\gamma, {\mathcal U}({\mathbb P}^\ell), E({\mathbb P}^\ell)\cap [\dot{A}^\ell_\gamma]^{\textstyle 2}): \ell\leq k\}\mbox{ is an ${\mathcal S}$--family of good graphs. ''}\] \end{lemma} \begin{proof} For a condition $q_0\in{\mathbb P}_\gamma$ we may find a stronger condition $q\in {\mathbb P}_\gamma$ with the following property \begin{enumerate} \item[$(*)_q$] for each $\xi\in I_\ell\cap\dom(q)$, $\ell\leq k$ there are $\bar{c}(q,\xi)$, $\bar{c}_0(q,\xi)$, $\bar{c}_1(q,\xi)$, and $U(q,\xi)$, $\bar{U}_0(q,\xi)$, $\bar{U}_1(q,\xi)$ (objects, not names) such that the condition $q{\restriction}\xi$ forces the following: \begin{itemize} \item $U(q,\xi)\in{\mathcal U}({\mathbb P}^\ell)$, $\dot{p}_\xi\in U(q,\xi)$, $q(\xi)= \bar{c}(q,\xi)\in{\mathbb C}^{\omega_1}({\mathbb P}^\ell,\dot{p}_\xi)$, \item $\bar{U}_0(q,\xi),\bar{U}_1(q,\xi):\dom(\bar{c}(q,\xi))\longrightarrow {\mathcal U}({\mathbb P}^\ell)$, $\bar{U}_0(q,\xi)(\varepsilon)\cap\bar{U}_1(q,\xi)( \varepsilon)=\emptyset$ for each $\varepsilon\in\dom(\bar{c}(q,\xi))$, \item for each $p'\in U(q,\xi)$ such that ${\mathbb C}({\mathbb P}^\ell,p')$ is defined: $\bar{c}(q,\xi),\bar{c}_0(q,\xi),\bar{c}_1(q,\xi)$ are in ${\mathbb C}^{\omega_1}( {\mathbb P}^\ell,p')$, the conditions $\bar{c}_0(q,\xi),\bar{c}_1(q,\xi)$ are stronger than $\bar{c}(q,\xi)$ and $\dom(\bar{c}_0(q,\xi))=\dom(\bar{c}_1(q,\xi))= \dom(\bar{c}(q,\xi))$, and if $i<2$, $\varepsilon\in\dom(\bar{c}(q,\xi))$, and $\dot{p}^*_\varepsilon$ is the ${\mathbb C}^{\omega_1}({\mathbb P}^\ell,p')$--name for the $\varepsilon^{\rm th}$ generic real (i.e., $\bigcup\{r:(\alpha,r)\in\Gamma_{{\mathbb C}^{\omega_1}( {\mathbb P}^\ell,p')}\}$) then \[\bar{c}_i(q,\xi)\forces_{{\mathbb C}^{\omega_1}({\mathbb P},p')}\mbox{`` } \dot{p}^*_\varepsilon\in \bar{U}_i(q,\xi)(\varepsilon)\mbox{ '',}\] \item for each $\varepsilon\in\dom(\bar{c}(q,\xi))$ and $q_0\in\bar{U}_0(q, \xi)(\varepsilon)$, $q_1\in\bar{U}_1(q,\xi)(\varepsilon)$ we have \[(\forall\varepsilon_0,\varepsilon_1<\omega_1)(\{(q_0,\varepsilon_0),(q_1, \varepsilon_1)\}\notin E({\mathbb P}^\ell)).\] \end{itemize} \end{enumerate} [Why? Just apply \ref{oneforce} (and remember that supports are finite).] >From now on we will restrict ourselves to conditions $q\in{\mathbb P}_\gamma$ with the property $(*)_q$ (what is allowed as they are dense in ${\mathbb P}_\gamma$). So we will assume that for each condition $q$ under considerations and $\xi\in I_\ell\cap\dom(q)$, $\ell\leq k$, the objects (not names!) $\bar{c}(q,\xi)$, $\bar{c}_0(q,\xi)$, $\bar{c}_1(q,\xi)$, $U(q,\xi)$, $\bar{U}_0(q,\xi)$, $\bar{U}_1(q,\xi)$ are defined and have the respective properties. Note that, in $\bV^{{\mathbb P}_\gamma}$, if $\ell\leq k$, $\xi_0,\xi_\ell\in I_\ell$, $(q,\alpha_0)\in \dot{Z}^{{\mathbb P}^\ell}_{\dot{p}^{\xi_0}}$ and $(q, \alpha_1)\in\dot{Z}^{{\mathbb P}^\ell}_{\dot{p}^{\xi_1}}$ then $\alpha_0=\alpha_1$ and $\xi_0=\xi_1$ (remember \ref{oneforce}). Therefore, we may label elements of $\dot{A}^\ell_\gamma$ by pairs from $I_\ell\times\omega_1$ and allow ourselves small abuse of notation identifying $(\xi,\alpha)\in I_\ell\times\omega_1$ with the respective $(q,\alpha)\in \dot{Z}^{{\mathbb P}^\ell}_{\dot{p}^\xi}$. Next let $E_\ell=E({\mathbb P}^\ell)\cap [\dot{A}^\ell_\gamma]^{\textstyle 2}$. Now, suppose that some condition $q'\in{\mathbb P}_\gamma$ forces that \[\{(\dot{A}^\ell_\gamma, {\mathcal U}({\mathbb P}^\ell),E_\ell): \ell\leq k\}\mbox{ is not an ${\mathcal S}$--family.}\] Then we may find a condition $q\in{\mathbb P}_\gamma$, an integer $m<\omega$, $\ell_0,\ldots,\ell_{m-1}\leq k$ (not necessarily distinct) and ${\mathbb P}_\gamma$--names $\dot{\nu}_\alpha$ (for $\alpha<\omega_1$) of sequences of length $m$ such that \[\begin{array}{ll} q\forces_{{\mathbb P}_\gamma}&\mbox{`` }(\forall\alpha<\beta<\omega_1)( \dot{\nu}_\alpha\neq\dot{\nu}_\beta)\ \&\ (\forall\alpha<\omega_1)(\forall i< m)(\dot{\nu}_\alpha(i)\in I_{\ell_i}\times\omega_1)\\ \ &\ \ (\forall\alpha<\beta<\omega_1)(\exists i<m)(\{\dot{\nu}_\alpha(i), \dot{\nu}_\beta(i)\}\in E_{\ell_i})\mbox{ ''.} \end{array}\] For each $\alpha<\omega_1$ choose a condition $q_\alpha\in{\mathbb P}_\gamma$ (satisfying $(*)_{q_\alpha}$ and) stronger than $q$ and a sequence $\nu_\alpha\in\prod\limits_{i<m}(I_{\ell_i}\times\omega_1)$ such that $q_\alpha\forces\dot{\nu}_\alpha=\nu_\alpha$ and \[(\forall i<m)(\nu_\alpha(i)=(\xi,\varepsilon)\ \Rightarrow\ \xi\in\dom(q_\alpha)\ \&\ \varepsilon\in\dom(q_\alpha(\xi))).\] Now we consider two cases. \smallskip \noindent{\sc Case A:}\qquad $\cf(\gamma)\neq\omega_1$.\\ Then for some $\zeta<\gamma$, for uncountably many $\alpha<\omega_1$, $\dom(q_\alpha)\subseteq\zeta$. Let $G\subseteq{\mathbb P}_\gamma$ be a generic over $\bV$ and work in $\bV[G\cap{\mathbb P}_\zeta]$. Because of the ccc of ${\mathbb P}_\zeta$, the set $\{\alpha<\omega_1: q_\alpha\in G\cap{\mathbb P}_\zeta\}$ is uncountable, so we get an uncountable $m$--selector from $\{(\dot{A}^\ell_\zeta, {\mathcal U}({\mathbb P}^\ell), E({\mathbb P}^\ell)\cap [\dot{A}^\ell_\zeta]^{\textstyle 2})^{G\cap {\mathbb P}_\zeta}:\ell\leq k\}$ (in $\bV[G\cap{\mathbb P}_\zeta]$), contradicting the assumption $(\gamma)$. \medskip \noindent{\sc Case B:}\qquad $\cf(\gamma)=\omega_1$.\\ If for some $\zeta<\gamma$ the set $\{\alpha<\omega_1:\dom(q_\alpha) \subseteq \zeta\}$ is uncountable then we may repeat the arguments of Case A. So assume that $\{\alpha<\omega_1:\dom(q_\alpha)\subseteq\zeta\}$ is countable for each $\zeta<\gamma$. Applying ``standard cleaning procedure'' and passing to an uncountable subsequence (and possibly increasing our conditions) we may assume that $|\dom(q_\alpha)|=N$ for each $\alpha<\omega_1$ and, letting $\{\xi^\alpha_0,\ldots,\xi^\alpha_{N-1}\}$ be the increasing enumeration of $\dom(q_\alpha)$: \begin{enumerate} \item $\{\dom(q_\alpha):\alpha<\omega_1\}$ forms a $\Delta$--system with heart $u^*$, \item for some $n^*<N$ and $\zeta^*<\gamma$, we have $(\forall\alpha<\omega_1)(\forall j<n^*)(\xi^\alpha_j<\zeta^*)$ and $(\forall\alpha<\beta<\omega_1)(\zeta^*<\xi^\alpha_{n^*}\leq\xi^\alpha_{N-1} <\xi^\beta_{n^*})$ (so necessarily $u^*\subseteq\zeta^*$), \item $\sup\{\xi^\alpha_{n^*}:\alpha<\omega_1\}=\gamma$, \item $(\forall\alpha,\beta<\omega_1)(\forall\ell\leq k)(\forall j<N)( \xi^\alpha_j\in I_\ell\ \Leftrightarrow\ \xi^\beta_j\in I_\ell)$, \item if $\alpha,\beta<\omega_1$, $\ell\leq k$, $j<N$ and $\xi^\alpha_j\in I_\ell$ then $U(q_\alpha,\xi^\alpha_j)=U(q_\beta,\xi^\beta_j)$, $\bar{c}(q_\alpha,\xi^\alpha_j)\sim \bar{c}(q_\beta,\xi^\beta_j)$, $\bar{c}_0(q_\alpha,\xi^\alpha_j)\sim \bar{c}_0(q_\beta,\xi^\beta_j)$, $\bar{c}_1(q_\alpha,\xi^\alpha_j)\sim \bar{c}_1(q_\beta,\xi^\beta_j)$ (see \ref{isomorph}), and \begin{enumerate} \item[$(*)$] if $H:\dom(\bar{c}(q_\alpha,\xi^\alpha_j))\longrightarrow\dom (\bar{c}(q_\beta,\xi^\beta_j))$ is the order preserving bijection then $H$ is the identity on $\dom(\bar{c}(q_\alpha,\xi^\alpha_j))\cap\dom(\bar{c}( q_\beta,\xi^\beta_j))$ and for each $\varepsilon\in\bar{c}(q_\alpha, \xi^\alpha_j)$ \[\begin{array}{l} \bar{U}_0(q_\alpha,\xi^\alpha_j)(\varepsilon)=\bar{U}_0(q_\beta, \xi^\beta_j)(H(\varepsilon)),\quad \bar{U}_1(q_\alpha,\xi^\alpha_j)( \varepsilon)=\bar{U}_1(q_\beta,\xi^\beta_j)(H(\varepsilon))\quad\mbox{ and}\\ (\forall i<m)(\nu_\alpha(i)=(\xi^\alpha_j,\varepsilon)\ \Leftrightarrow\ \nu_\beta(i)=(\xi^\beta_j,H(\varepsilon))). \end{array}\] \end{enumerate} \end{enumerate} Let $w^*$ be the set of these $i<m$ that for some (equivalently: all) $\alpha<\omega_1$ we have $\nu_\alpha(i)\in\zeta^*\times\omega_1$. \begin{claim} \label{clx1} There are $q^*\in{\mathbb P}_{\zeta^*}$ and $\alpha<\beta<\omega_1$ such that $q^*$ is stronger than both $q_\alpha{\restriction}\zeta^*$ and $q_\beta{\restriction}\zeta^*$ and $(\forall i\in w^*)(\{\nu_\alpha(i),\nu_\beta(i)\}\notin E_{\ell_i})$. \end{claim} \begin{proof}[Proof of the claim] Let $G_{\zeta^*}\subseteq{\mathbb P}_{\zeta^*}$ be a generic filter over $\bV$. Work in $\bV[G_{\zeta^*}]$. By the ccc of ${\mathbb P}_{\zeta^*}$, the set $\{\alpha< \omega_1:q_\alpha{\restriction}\zeta^*\in G_{\zeta^*}\}$ is uncountable. Look at the sequence $\langle\nu_\alpha{\restriction} w^*:q_\alpha{\restriction}\zeta^*\in G_{\zeta^*} \rangle$. By assumption $(\gamma)$ of the lemma, it cannot be a $|w^*|$--selector, so there are $\alpha<\beta<\omega_1$ such that \[q_\alpha{\restriction}\zeta^*\quad\&\quad q_\beta{\restriction}\zeta^*\quad\&\quad (\forall i\in w^*)(\{\nu_\alpha(i),\nu_\beta(i)\}\notin E_{\ell_i}).\] Now, going back to $\bV$, we easily find a condition $q^*\in{\mathbb P}_{\zeta^*}$ such that $q^*,\alpha,\beta$ are as required. \end{proof} Let $q^*,\alpha,\beta$ be as guaranteed by \ref{clx1}. For $j<N$ such that $\xi^\alpha_j\in I_\ell$, $\ell\leq k$, let $H_j:\dom(\bar{c}(q_\alpha, \xi^\alpha_j))\longrightarrow\dom(\bar{c}(q_\beta,\xi^\beta_j))$ be the order preserving bijection (see clause (5) above). We define a condition $q^+\in{\mathbb P}_{\gamma}$ as follows:\\ $\dom(q^+)=\dom(q^*)\cup\dom(q_\alpha)\cup\dom(q_\beta)$ and \begin{itemize} \item $q^+{\restriction} \zeta^*=q^*$, \item if $\xi^\alpha_j\in I_\ell$, $n^*\leq j<N$, $\ell\leq k$ then \[q^+(\xi^\alpha_j)=\bar{c}_0(q_\alpha,\xi^\alpha_j)\quad\mbox{ and }\quad q^+(\xi^\beta_j)=\bar{c}_0(q_\beta,\xi^\beta_j),\] \item if $n^*\leq j<N$, $\xi^\alpha_j\notin\bigcup\limits_{\ell\leq k} I_\ell$ then $q^+(\xi^\alpha_j)=q_\alpha(\xi^\alpha_j)$, $q^+(\xi^\beta_j)= q_\beta(\xi^\beta_j)$. \end{itemize} It should be clear that $q^+\in{\mathbb P}_\gamma$ is a condition stronger than both $q_\alpha$ and $q_\beta$. If $i<m$ and $\varepsilon<\omega_1$ then \[\nu_\alpha(i)=(\xi^\alpha_j,\varepsilon)\ \Leftrightarrow\ \nu_\beta(i)= (\xi^\beta_j,H_j(\varepsilon)).\] If $i\in m\setminus w^*$, $\nu_\alpha(i)=(\xi^\alpha_j,\varepsilon)$, $n^*\leq j<N$ and $\dot{p}^*_{\varepsilon,\xi^\alpha_j}$, $\dot{p}^*_{H(\varepsilon),\xi^\beta_j}$ are the names for $\varepsilon^{\rm th}$ ($H(\varepsilon)^{\rm th}$ respectively) generic reals added by $\dot{{\mathbb Q}}_{\xi^\alpha_j}$ ($\dot{{\mathbb Q}}_{\xi^\beta_j}$, resp.) then \[\begin{array}{ll} q^+\forces_{{\mathbb P}_\gamma}&\mbox{`` }\dot{p}^*_{\varepsilon,\xi^\alpha_j}\in \bar{U}_0(q_\alpha,\xi^\alpha_j)(\varepsilon)=\bar{U}_0(q_\beta,\xi^\beta_j) (H(\varepsilon))\quad\mbox{ and}\\ \ &\ \ \dot{p}^*_{H(\varepsilon),\xi^\beta_j}\in\bar{U}_1(q_\beta, \xi^\beta_j)(H(\varepsilon))=\bar{U}_1(q_\alpha,\xi^\alpha_j)(\varepsilon) \mbox{ ''.} \end{array}\] If $i\in w^*$ then look at the choice of $q^*,\alpha,\beta$ (see \ref{clx1}). Putting everything together we conclude that \[q^+\forces\mbox{`` }(\forall i<m)(\{\dot{\nu}_\alpha(i),\dot{\nu}_\beta(i) \}\notin E_{\ell_i})\mbox{ ''},\] a contradiction \end{proof} \section{Proof of Theorem \ref{main}} Let $\kappa$ be regular cardinal such that $\kappa=\kappa^{<\kappa}\geq \aleph_2$. By induction on $\xi\leq\kappa$ we build a finite support iteration $\langle{\mathbb P}_\xi,\dot{{\mathbb Q}}_\xi:\xi<\kappa\rangle$ and sequences $\langle\dot{{\mathbb P}}^\xi,I_\xi:\xi<\kappa\rangle$, $\langle\dot{A}^\xi_\zeta: \xi<\zeta\leq\kappa\rangle$ and $\langle\dot{p}^\zeta:\zeta\in I_\xi\rangle$ such that for each $\xi,\xi_0,\xi_1<\kappa$ \begin{enumerate} \item[(i)] $\dot{{\mathbb Q}}_\xi$ is a ${\mathbb P}_\xi$--name for a ccc forcing notion on a bounded subset of $\kappa$, \item[(ii)] $I_\xi\in [\{2\cdot\alpha:\xi<\alpha<\kappa\}]^{\textstyle \kappa}$, $\dot{{\mathbb P}}^\xi$ is a ${\mathbb P}_\xi$--name for an element of ${\mathcal K}$, and if $\xi_0\neq\xi_1$ then $I_{\xi_0}\cap I_{\xi_1}=\emptyset$, \item[(iii)] for $\zeta\in I_\xi$, $\dot{p}^\zeta$ is a ${\mathbb P}_\zeta$--name for a condition in $\dot{{\mathbb P}}^\xi$ for which ${\mathbb C}(\dot{{\mathbb P}}^\xi,\dot{p}^\zeta)$ is defined, \item[(iv)] if $\zeta\in I_\xi$ then $\dot{{\mathbb Q}}_\zeta$ is (equivalent to) ${\mathbb C}^{\omega_1}(\dot{{\mathbb P}}^\xi,\dot{p}^\zeta)$, \item[(v)] if $\xi<\zeta\leq\kappa$ then $\dot{A}^\xi_\zeta$ is the ${\mathbb P}_\zeta$--name for the set $\bigcup\{\dot{Z}^{\dot{{\mathbb P}}^\xi}_{ \dot{p}^\varepsilon}:\varepsilon\in I_\xi\cap\zeta\}\subseteq\dot{{\mathbb P}}^\xi \times\omega_1$ (where $\dot{Z}^{\dot{{\mathbb P}}^\xi}_{\dot{p}^\varepsilon}$ is the generic object added by $\dot{{\mathbb Q}}_\zeta$; compare \ref{maintech}), \item[(vi)] for each $\zeta\leq\kappa$ \[\forces_{{\mathbb P}_\zeta}\mbox{`` }\{(\dot{A}^\varepsilon_\zeta,{\mathcal U}( \dot{{\mathbb P}}^\varepsilon),E(\dot{{\mathbb P}}^\varepsilon)\cap [ \dot{A}^\varepsilon_\zeta]^{\textstyle 2}): \varepsilon<\zeta\}\mbox{ is an ${\mathcal S}$--family of good graphs '',}\] \item[(vii)] if $\dot{{\mathbb Q}}$ is a ${\mathbb P}_\kappa$--name for a ccc forcing notion on a bounded subset of $\kappa$ then \[|\{\zeta<\kappa:\ \ \forces_{{\mathbb P}_\zeta}\mbox{`` }\dot{{\mathbb Q}}_\zeta=\dot{{\mathbb Q}} \mbox{ ''}\}|=\kappa,\] \item[(viii)] if $\dot{{\mathbb P}}$ is a ${\mathbb P}_\kappa$--name for an element of ${\mathcal K}$ then then for some $\varepsilon<\kappa$ we have \[\forces_{{\mathbb P}_\varepsilon}\mbox{`` }\dot{{\mathbb P}}^\varepsilon=\dot{{\mathbb P}}\mbox{ ''\ \ and\ \ }\forces_{{\mathbb P}_\kappa}\mbox{`` the set }\{\dot{p}^\zeta:\zeta \in I_\varepsilon\} \mbox{ is dense in $\dot{{\mathbb P}}$ ''.}\] \end{enumerate} We use the standard bookkeeping arguments to choose the lists $\langle \dot{{\mathbb P}}^\xi,I_\xi:\xi<\kappa\rangle$, $\langle\dot{p}^\zeta:\zeta\in I_\xi \rangle$ so that clauses (ii), (iii) and (viii) are satisfied. Similarly we choose a list $\langle\dot{{\mathbb Q}}'_\xi:\xi<\kappa\rangle$ of all ${\mathbb P}_\kappa$--names for partial orders on bounded subsets of $\kappa$ so that each name appears $\kappa$ many times in the list, and $\dot{{\mathbb Q}}'_\xi$ is a ${\mathbb P}_\xi$--name (for $\xi<\kappa$) (this list will be used to take care of clauses (i), (vii)). Now we have to be more specific. So suppose that for some $\xi<\kappa$ we have already defined the iteration $\langle {\mathbb P}_\zeta,\dot{{\mathbb Q}}_\zeta:\zeta< \xi\rangle$. If $\xi$ is a limit ordinal, before we go further we should argue that the clause (vi) is satisfied by the limit ${\mathbb P}_\xi$, i.e., \[\forces_{{\mathbb P}_\xi}\mbox{`` }{\mathcal F}_\xi\stackrel{\rm def}{=}\{(\dot{A }^\varepsilon_\zeta,{\mathcal U}(\dot{{\mathbb P}}^\varepsilon),E(\dot{{\mathbb P}}^\varepsilon)\cap [\dot{A}^\varepsilon_\zeta]^{\textstyle 2}): \varepsilon<\zeta\}\mbox{ is an ${\mathcal S}$--family of good graphs ''.}\] But this is immediate by \ref{maintech} --- if a problem occurs than it is caused by a finite subfamily of ${\mathcal F}$ and we may assume that the respective forcing notions ${\mathbb P}^{\varepsilon_\ell}$ are from the ground model. Suppose $\xi=2\cdot\alpha+1$. Then we look at the ${\mathbb P}_\alpha$--name ${\mathbb Q}'_\alpha$ and we ask if, in $\bV^{{\mathbb P}_\xi}$, it is a ccc forcing notion. If not that we let ${\mathbb Q}_\xi$ be the Cohen real forcing. If yes, then we we ask if (in $\bV^{{\mathbb P}_\xi}$) it forces that ${\mathcal F}_\xi$ remains an ${\mathcal S}$--family. If again yes, then we let $\dot{{\mathbb Q}}_\xi$ be $\dot{{\mathbb Q}}'_\alpha$; otherwise $\dot{{\mathbb Q}}_\xi={\mathbb P}^{{\mathcal F}_\xi}$ (see \ref{notccc}). In any case we are sure that the relevant instances of clauses (i)--(viii) are satisfied (remember \ref{conclusion}). Assume now that $\xi=2\cdot\alpha\in I_\zeta$, $\zeta<\kappa$. Then clause (iv) determines $\dot{{\mathbb Q}}_\xi$. We should show that the clause (vi) holds true. Suppose that we may find a condition $q\in{\mathbb P}_{\xi+1}$, an integer $m<\omega$, $\varepsilon_0,\ldots,\varepsilon_{m-1}\leq\xi$ and ${\mathbb P}_{\xi+1}$--names $\dot{\nu}_\alpha$ (for $\alpha<\omega_1$) of sequences of length $m$ such that \[\begin{array}{ll} q\forces_{{\mathbb P}_{\xi+1}}&\mbox{`` }(\forall\alpha<\beta<\omega_1)( \dot{\nu}_\alpha\neq\dot{\nu}_\beta)\ \&\ (\forall\alpha<\omega_1)(\forall i< m)(\dot{\nu}_\alpha(i)\in \dot{A}^{\varepsilon_i}_{\xi+1})\\ \ &\ \ (\forall\alpha<\beta<\omega_1)(\exists i<m)(\{\dot{\nu}_\alpha(i), \dot{\nu}_\beta(i)\}\in E(\dot{{\mathbb P}}^{\varepsilon_i}))\mbox{ ''.} \end{array}\] We may additionally demand that for some $k<m$ we have \[\begin{array}{ll} q\forces_{{\mathbb P}_{\xi+1}}&\mbox{`` }(\forall\alpha<\omega_1)(\forall i<k) (\dot{\nu}_\alpha(i)\in \dot{A}^{\varepsilon_i}_{\xi})\quad\mbox{ and}\\ \ &\ \ (\forall\alpha<\omega_1)(\forall i\in [k,m))(\varepsilon_i=\xi\ \&\ \dot{\nu}_\alpha(i)\in \dot{Z}^{\dot{{\mathbb P}}^\zeta}_{ \dot{p}^\xi})\mbox{ ''.} \end{array}\] \begin{claim} \label{clx2} Suppose that ${\mathbb P}\in{\mathcal K}$, $p\in{\mathbb P}$ (and ${\mathbb C}({\mathbb P},p)$ is defined). Then \[\forces_{{\mathbb C}^{\omega_1}({\mathbb P},p)}\mbox{`` }(\forall s_0,s_1\in \dot{Z}^{\mathbb P}_p)(\{s_0,s_1\}\notin E({\mathbb P}))\mbox{ ''.}\] \end{claim} \begin{proof}[Proof of the claim] Like \ref{oneforce}. \end{proof} It follows from \ref{clx2} that \[q\forces_{{\mathbb P}_{\xi+1}}\mbox{`` }(\forall\alpha<\beta<\omega_1)(\exists i< k)(\{\dot{\nu}_\alpha(i),\dot{\nu}_\beta(i)\}\in E(\dot{{\mathbb P}}^{\varepsilon_i}))\mbox{ ''.}\] But, by \ref{conclusion}, we have \[\forces_{{\mathbb P}_{\xi+1}}\mbox{`` ${\mathcal F}_\xi$ is an ${\mathcal S}$--family of good graphs '',}\] so we get an immediate contradiction. Finally if $\xi=2\cdot\alpha\notin\bigcup\limits_{\zeta<\kappa} I_\zeta$ then we let ${\mathbb Q}_\xi$ be the Cohen real forcing (again all clauses are preserved). The construction is complete. We claim that the limit forcing notion ${\mathbb P}_\kappa$ is as required in \ref{main}. Clearly it satisfies the ccc and (a dense subset of it) is of size $\kappa$. Clause (vii) guarantees that $\forces_{{\mathbb P}_\kappa}$`` $\con=\kappa\ \&\ {\bf MA}$ ''. It follows from the clause (vi) and \ref{maintech} that \[\forces_{{\mathbb P}_\xi}\mbox{`` }{\mathcal F}_\kappa\stackrel{\rm def}{=}\{(\dot{A }^\varepsilon_\zeta,{\mathcal U}(\dot{{\mathbb P}}^\varepsilon),E(\dot{{\mathbb P}}^\varepsilon)\cap [\dot{A}^\varepsilon_\zeta]^{\textstyle 2}): \varepsilon<\kappa\}\mbox{ is an ${\mathcal S}$--family of good graphs ''.}\] Clauses (v)+(viii) (and the definition of $E(\dot{{\mathbb P}}^\varepsilon)$) imply that for each $\varepsilon<\kappa$ \[\forces_{{\mathbb P}_\kappa}\mbox{`` }(\dot{A}^\varepsilon_\zeta,{\mathcal U}(\dot{{\mathbb P} }^\varepsilon),E(\dot{{\mathbb P}}^\varepsilon)\cap [\dot{A}^\varepsilon_\zeta]^{ \textstyle 2})\mbox{ is densely representably by $\bot_{\dot{{\mathbb P}}^\varepsilon}$ ''}.\] Consequently, by \ref{notFAP} and clause (viii) we get \[\forces_{{\mathbb P}_\kappa}\mbox{`` }(\forall{\mathbb P}\in{\mathcal K})(\neg{\bf FA}_{\aleph_1}({\mathbb P})) \mbox{ '',}\] finishing the proof of \ref{main}. \begin{corollary} It is consistent with ${\bf MA}+\neg{\rm CH}$ that any forcing notion ${\mathbb P}\in{\mathcal K}$ collapses $\con$ to $\omega_1$ (and thus is not $\omega_1$--proper). \end{corollary} \section{Open problems} The model constructed in the previous section provides $(\forall{\mathbb P}\in{\mathcal K})(\neg{\bf FA}_{\aleph_1}({\mathbb P}))$ by dealing with each forcing ${\mathbb P}\in{\mathcal K}$ separately. We would like to have one common reason for $\neg{\bf FA}_{\aleph_1}({\mathbb P})$ for all forcing notions ${\mathbb P}\in{\mathcal K}$, i.e., a combinatorial principle ${\mathcal P}$ which is consistent with ${\bf MA}+\neg {\rm CH}$ and which implies $(\forall{\mathbb P}\in{\mathcal K})(\neg{\bf FA}_{\aleph_1}({\mathbb P}))$. A possible candidate for a principle like that was already pointed in \cite[\S 2]{CRSW93}. As we stated in the Introduction, our method is a slight generalization of that of Steprans. Steprans' method in turn was based on the proof of Abraham, Rubin and Shelah \cite{ARSh:153} that it is consistent with ${\bf MA}+\neg {\rm CH}$ that there are two non-isomorphic $\aleph_1$--dense sets of reals. In the latter proof, a 2--entangled set of reals was used. This leads us to the following question \begin{problem} [Compare {\cite[Question 2.4]{CRSW93}}] Does the existence of a 2--entangled set of reals of size $\aleph_1$ imply $(\forall{\mathbb P}\in{\mathcal K})(\neg{\bf FA}_{\aleph_1}({\mathbb P}))$? \end{problem} If one tries to repeat the proof of \cite[Theorem 2.1]{JMSh:372} for elements of ${\mathcal K}$ then one gets into some problems in cases 1,3 of Definition \ref{classK}. Possible reason for it is that a proof as in \cite[Theorem 2.1]{JMSh:372} would give a property that seems to be stronger than $\neg{\bf FA}_{\aleph_1}({\mathbb P})$. \begin{definition} Let ${\mathbb P}$ be a forcing notion of size $\con$, $\bar{p}=\langle p_i: i<\con \rangle\subseteq{\mathbb P}$. We say that $\bar{p}$ is an JMSh--sequence if \begin{enumerate} \item[$(\oplus)_{\rm JMSh}$] given $\langle F_\alpha:\alpha<\omega_1\rangle$ pairwise disjoint finite subsets of $\con$, there exist $\alpha<\beta< \omega_1$ such that \[(\forall i\in F_\alpha)(\forall j\in F_\beta)(p_i\, \bot_{{\mathbb P}}\; p_j).\] \end{enumerate} \end{definition} \begin{proposition} \label{JMShseq} Suppose that ${\mathbb P}$ is a forcing notion of size $\con$ such that \begin{enumerate} \item above any condition in ${\mathbb P}$, there is an antichain (in ${\mathbb P}$) of size $\con$, and \item there is an JMSh--sequence $\bar{p}\subseteq{\mathbb P}$ which is dense in ${\mathbb P}$. \end{enumerate} Then $\neg{\bf FA}_{\aleph_1}({\mathbb P})$. \end{proposition} \begin{problem} \begin{enumerate} \item Is the existence of dense JMSh--sequences in ${\mathbb P}$ really stronger than $\neg{\bf FA}_{\aleph_1}({\mathbb P})$ (for ${\mathbb P}$ of size $\con$ satisfying the assumption \ref{JMShseq}(1)) ? \item Is it consistent with ${\bf MA}+\neg{\rm CH}$ that for each ${\mathbb P}\in{\mathcal K}$ there is a dense JMSh--sequence in ${\mathbb P}$? \end{enumerate} \end{problem} On the other hand, Judah, Miller and Shelah \cite{JMSh:372} and Goldstern, Johnson and Spinas \cite{GJS94} showed that ${\bf MA}_{\omega_1}(\mbox{ccc} )$ implies the forcing axiom for the Miller and Laver forcing notions. This gives a strong expectation that ${\bf MA}_{\omega_1}(\mbox{ccc})$ implies forcing axioms for most of forcing notions (with norms) adding unbounded reals. Brendle \cite[Proposition 5.1]{Br96a} showed that ${\bf MA}$ implies that the Laver forcing, the Mathias forcing, the Miller forcing and the Blass-Shelah forcing are $\alpha$--proper for all $\alpha<\con$. Again, one expects that this could be generalized further. \begin{problem} Let ${\mathcal K}^\bot$ be the class of the forcing notions of \cite{RoSh:470} which are not in ${\mathcal K}$ for nontrivial reasons. \begin{enumerate} \item Does ${\bf MA}+\neg{\rm CH}$ imply ${\bf FA}_{\aleph_1}({\mathbb P})$ for all ${\mathbb P}\in {\mathcal K}^\bot$ ? \item Does ${\bf MA}+\neg{\rm CH}$ imply that all ${\mathbb P}\in{\mathcal K}^\bot$ are $\alpha$--proper (for all $\alpha<\con$) ? \end{enumerate} \end{problem} Finally, possible further generalizations of the present paper could go into the direction of nep/snep forcing notions of Shelah \cite{Sh:630}, \cite{Sh:669}. \begin{problem} \begin{enumerate} \item Is ${\bf MA}+\neg{\rm CH}$ consistent with the failure of ${\bf FA}_{\aleph_1}({\mathbb P})$ for all snep $\baire$--bounding forcing notions ${\mathbb P}$ which do not have ccc above any condition? \item Does ${\bf MA}+\neg{\rm CH}$ imply ${\bf FA}_{\aleph_1}({\mathbb P})$ for all snep forcing notions ${\mathbb P}$ adding unbounded reals? \end{enumerate} \end{problem}
\section{Introduction} In recent years, there has been a renaissance of interest in diffractive scattering. These diffractive processes are described by the Regge theory in terms of the Pomeron ($I\!\! P$) exchange\cite{pomeron}. The Pomeron carries quantum numbers of the vacuum, so it is a colorless entity in QCD language, which may lead to the ``rapidity gap" events in experiments. However, the nature of Pomeron and its reaction with hadrons remain a mystery. For a long time it had been understood that the dynamics of the ``soft pomeron'' is deeply tied to confinement. However, it has been realized now that how much can be learned about QCD from the wide variety of small-$x$ and hard diffractive processes, which are now under study experimentally. In Refs.\cite{th1,th2}, the diffractive $J/\psi$ and $\Upsilon$ production cross section have been formulated in photoproduction processes and in DIS processes in perturbative QCD. In the framework of perturbative QCD the Pomeron is represented by a pair of gluon in the color-singlet sate. This two-gluon exchange model can successfully describe the experimental results from HERA\cite{hera-ex}. On the other hand, as we know that there exist nonfactorization effects in the hard diffractive processes at hadron colliders \cite{preqcd,collins,soper,tev}. First, there is the so-called spectator effect\cite{soper}, which can change the probability of the diffractive hadron emerging from collisions intact. Practically, a suppression factor (or survive factor) ``$S_F$" is used to describe this effect\cite{survive}. Obviously, this suppression factor can not be calculated in perturbative QCD, which is now viewed as a nonperturbative parameter. Typically, the suppression factor $S_F$ is determined to be about $0.1$ at the energy scale of the Fermilab Tevatron\cite{tev}. Another nonfactorization effect discussed in literature is associated with the coherent diffractive processes at hadron colliders\cite{collins}, in which the whole Pomeron is induced in the hard scattering. It is proved in \cite{collins} that the existence of the leading twist coherent diffractive processes is associated with a breakdown of the QCD factorization theorem. Based on the success of the two-gluon exchange parametrization of the Pomeron model in the description of the diffractive photoproduction processes at $ep$ colliders\cite{th1,th2,hera-ex}, we may extend the applications of this model to calculate the diffractive processes at hadron colliders in perturbative QCD. Under this context, the Pomeron represented by a color-singlet two-gluon system emits from one hadron and interacts with another hadron in hard process, in which the two gluons are both involved (as shown in Fig.~1). Therefore, these processes calculated in the two-gluon exchange model are just belong to the coherent diffractive processes in hadron collisions. Another important feature of the calculations of the diffractive processes in this model recently demonstrated is the sensitivity to the off-diagonal parton distribution function in the proton\cite{offd}. Using this two-gluon exchange model, we have calculated the diffractive $J/\psi$ production \cite{psi}, quark jet production\cite{charm,quark}, massive muon pair and $W$ boson productions\cite{dy}, and direct photon production\cite{photon} in hadron collisions. In this paper, we will further calculate the gluon jet production at large transverse momentum in the coherent diffractive processes at hadron colliders by using the two-gluon exchange model. In the calculations of Refs.\cite{psi,charm,dy}, there always is a large mass scale associated with the production process. That is $M_\psi$ for $J/\psi$ production, $m_c$ for the charm jet production, $M^2$ for the massive muon production ($M^2$ is the invariant mass of the muon pair) and $M_W^2$ for $W$ boson production. However, in the gluon jet production process as well as the light quark jet production process, there is no large mass scale. So, for these processes, the large transverse momentum is needed to guarantee the application of the perturbative QCD. Furthermore, in \cite{quark} we show that the light quark jet production in the two-gluon exchange model has a distinctive feature that there is no contribution from the small $l_T^2$ region ($l_T^2<k_T^2$) in the integration of the amplitude over $l_T^2$. (The similar behavior has also been found for the diffractive light quark photoproduction process\cite{zaka}.) So, the expansion (in terms of $l_T^2/M_X^2$) method used in Refs.\cite{psi,charm,dy} can not be applied to the calculations of gluon jet production. In the following calculations, we will employ the helicity amplitude method to calculate the amplitude of the diffractive gluon jet production in hadron collisions. We will show that the production cross section is related to the differential (off-diagonal) gluon distribution function in the proton as that in the diffractive light quark jet production process\cite{quark}. (On the other hand, we note that the cross sections of the processes calculated in Refs.\cite{psi,charm,dy} are related to the integrated gluon distribution function in the proton). Diffractive gluon jet production can come from two types partonic processes: one is the quark initiated process (Fig.2), and the other is the gluon initiated process (Fig.3). The diffractive production of heavy quark jet at hadron colliders has also been studied by using the two-gluon exchange model in Ref.\cite{levin}. However, their calculation method is very different from ours \footnote{For detailed discussions and comments, please see \cite{charm}}. In their calculations, they separated their diagrams into two parts, and called one part the coherent diffractive contribution to the heavy quark production. However, this separation can not guarantee the gauge invariance\cite{charm}. In our approach, we follow the definition of Ref.\cite{collins}, i.e., we call the process in which the whole Pomeron participants in the hard scattering process as the coherent diffractive process. Under this definition, all of the diagrams plotted in Fig.2 and Fig.3 for the partonic processes contribute to the coherent diffractive production. The rest of the paper is organized as follows. In Sec.II, we will give the cross section formula for the partonic process in the leading order of perturbative QCD. In this section we employ the helicity amplitude method to calculate two partonic processes, $qp\rightarrow qgp$ and $gp\rightarrow ggp$. In Sec.III, we estimate the production rate of diffractive gluon jet at the Fermilab Tevatron by approximating the off-diagonal gluon distribution function by the usual diagonal gluon distribution function in the proton. We also compare the contributions from different partonic processes to the diffractive dijet production at the Tevatron. And the conclusions will be given in Sec.IV. \section{ The cross section formula for the partonic process} \subsection{$qp\rightarrow qgp$ process} For the partonic process $qp\rightarrow qgp$, in the leading order of perturbative QCD, there are nine diagrams shown in Fig.2. The two-gluon system coupled to the proton (antiproton) in Fig.2 is in a color-singlet state, which characterizes the diffractive processes in perturbative QCD. Due to the positive signature of these diagrams (color-singlet exchange), we know that the real part of the amplitude cancels out in the leading logarithmic approximation. To get the imaginary part of the amplitude, we must calculate the discontinuity represented by the crosses in each diagram of Fig.2. The first four diagrams of Fig.2 are the same as those calculated in the diffractive direct photon production process at hadron colliders\cite{photon}. But, due to the existence of gluon-gluon interaction vertex in QCD, in the partonic process $qp\rightarrow qgp$, there are additional five diagrams (Fig.2(5)-(9)). These five diagrams are needed for a complete calculation in this order of QCD. In our calculations, we express the formulas in terms of the Sudakov variables. That is, every four-momenta $k_i$ are decomposed as, \begin{equation} k_i=\alpha_i q+\beta_i p+\vec{k}_{iT}, \end{equation} where $q$ and $p$ are the momenta of the incident quark and the proton, $q^2=0$, $p^2=0$, and $2p\cdot q=W^2=s$. Here $s$ is the c.m. energy of the quark-proton system, i.e., the invariant mass of the partonic process $qp\rightarrow qg p$. $\alpha_i$ and $\beta_i$ are the momentum fractions of $q$ and $p$ respectively. $k_{iT}$ is the transverse momentum, which satisfies \begin{equation} k_{iT}\cdot q=0,~~~ k_{iT}\cdot p=0. \end{equation} All of the Sudakov variables for every momentum are determined by using the on-shell conditions of the momenta represented by the external lines and the crossed lines in the diagram. The calculations of these Sudakov variables are similar to those in the diffractive light quark jet production process $gp\rightarrow q\bar q p$\cite{quark}, and we can get the Sudakov variables of every momentum for the process $qp\rightarrow qgp$ from the relevant formulas of \cite{quark}. In the following, we list all of the Sudakov variables for the diffractive process $qp\rightarrow qgp$. For the momentum $u$, we have \begin{equation} \alpha_u=0,~~\beta_u=x_{I\! P}=\frac{M_X^2}{s},~~u_T^2=t=0, \end{equation} where $M_X^2$ is the invariant mass squared of the diffractive final state including the light quark and antiquark jets. For the high energy diffractive process, we know that $M_x^2\ll s$, so we have $\beta _u$ ($x_{I\! P}$) as a small parameter. For the momentum $k$, \begin{equation} \label{ak} \alpha_k(1+\alpha_k)=-\frac{k_T^2}{M_X^2},~~\beta_k=-\alpha_k\beta_u, \end{equation} where $k_T$ is the transverse momentum of the out going quark jet. For the loop mentum $l$, because the results for $\beta_l$ are not the same for the nine diagrams of Fig.2, we get its value from the formula of Ref.\cite{quark} for the relevant diagram. The results are \begin{eqnarray} \nonumber \alpha_l&=&-\frac{l_T^2}{s},\\ \nonumber \beta_l&=&\frac{2(k_T,l_T)-l_T^2}{\alpha_ks},~~~{\rm for~Diag.}1,~2,~6,\\ \nonumber &=&\frac{2(k_T,l_T)+l_T^2}{(1+\alpha_k)s},~~~{\rm for~Diag.}5,~7,~8,\\ &=&-\frac{M_X^2-l_T^2}{s},~~~~~~~{\rm for~Diag.}3,~4,~9, \end{eqnarray} where $(k_T,l_T)$ is the 2-dimensional product of the transverse vectors $\vec{k}_T$ and $\vec{l}_T$. Using these Sudakov variables, we can give the cross section formula for the partonic process $qp\rightarrow qgp$ as, \begin{equation} \label{xs} \frac{d\hat{\sigma}(qp\rightarrow qgp)}{dt}|_{t=0}=\frac{dM_X^2d^2k_Td\alpha_k}{16\pi s^216\pi^3M_X^2} \delta(\alpha_k(1+\alpha_k)+\frac{k_T^2}{M_X^2})\sum \overline{|{\cal A}|}^2, \end{equation} where ${\cal A}$ is the amplitude of the process $qp\rightarrow qgp$. We know that the real part of the amplitude ${\cal A}$ is zero, and the imaginary part of the amplitude ${\cal A}(qp\rightarrow qgp)$ for each diagram of Fig.2 has the following general form, \begin{equation} \label{ima} {\rm Im}{\cal A}=C_F(T_{ij}^a)\int \frac{d^2l_T}{(l_T^2)^2}F\times\bar u _i(u-k)\Gamma_\mu u_j(q), \end{equation} where $C_F$ is the color factor for each diagram. $a$ is the color index of the incident gluon. $\Gamma_\mu$ represents some $\gamma$ matrices including one propagator. $F$ in the integral is defined as \begin{equation} \label{feq} F=\frac{3}{2s}g_s^3f(x',x^{\prime\prime};l_T^2), \end{equation} where \begin{equation} \label{offd1} f(x',x^{\prime\prime};l_T^2)=\frac{\partial G(x',x^{\prime\prime};l_T^2)}{\partial {\rm ln} l_T^2}, \end{equation} where the function $G(x',x^{\prime\prime};k_T^2)$ is the so-called off-diagonal gluon distribution function\cite{offd}. Here, $x'$ and $x^{\prime\prime}$ are the momentum fractions of the proton carried by the two gluons. It is expected that at small $x$, there is no big difference between the off-diagonal and the usual diagonal gluon densities\cite{off-diag}. So, in the following calculations, we estimate the production rate by approximating the off-diagonal gluon density by the usual diagonal gluon density, $G(x',x^{\prime\prime};Q^2)\approx xg(x,Q^2)$, where $x=x_{I\!\! P}=M_X^2/s$. In \cite{photon}, we calculate the cross section for the diffractive photon production process $qp\rightarrow \gamma qp$ by directly squaring the partonic process amplitude. However, in the calculations here for the partonic process $qp\rightarrow qgp$ because there are additional five diagrams contribution, it is not convenient to directly square the amplitude. Following Ref.\cite{quark}, we calculate the amplitude by employing the {\it helicity amplitude} method\cite{ham,wu}. Furthermore, we will show that by using the helicity amplitude method we can reproduce the cross section formula for the diffractive photon production process\cite{photon}. For the massless quark spinors, we define \begin{equation} u_\pm(p)=\frac{1}{\sqrt{2}}(1\pm\gamma_5)u(p). \end{equation} For the polarization vector of the outgoing gluon (its momentum is $k+q$), following the method of Ref.\cite{wu}, we find that it is convenient to choose \begin{equation} \label{epq} \not\! e^{(\pm)}=N_e[(\not\! k+\not\! q)\not\! q\not\! p(1\mp\gamma_5)+ \not\! p\not\! q(\not\! k+\not\! q)(1\pm\gamma_5)]. \end{equation} The normalization factor $N_e$ equals to \begin{equation} \label{ene} N_e=\frac{1}{s\sqrt{2k_T^2}}. \end{equation} With this definition (\ref{epq}), we can easily get the scalar products between the four-momenta and the polarization vector $e$ as \begin{equation} \label{eprc} e\cdot p=0,~~e\cdot q=N_e\frac{k_T^2s}{1+\alpha_k},~~e\cdot k_T=-N_ek_T^2s,~~e\cdot l_T=-N_e(k_T,l_T)s. \end{equation} The helicity amplitudes for the processes in which the polarized Dirac particles are involved have the following general forms\cite{ham}, \begin{equation} \label{ham} \bar u_\pm(p_f)Qu_\pm(p_i)=\frac{Tr[Q\not\! p_i\not\! n\not\! p_f(1\mp\gamma_5)]} {4\sqrt{(n\cdot p_i)(n\cdot p_f)}}, \end{equation} where $n$ is an arbitrary massless 4-vector, which is set to be $n=p$ in the following calculations. Using this formula (\ref{ham}), the calculations of the helicity amplitude ${\cal A}(\lambda_1,\lambda_2,\lambda_3)$ for the diffractive process $qp\rightarrow qgp$ is straightforward. Here $\lambda_1$ represents the helicity of the incident quark; $\lambda_2$ and $\lambda_3$ represent the helicities of the outgoing gluon and quark respectively. In our calculations, we only take the leading order contributions, and neglect the higher order contributions which are proportional to $\beta_u=\frac{M_X^2}{s}$ because in the high energy diffractive processes we have $\beta_u\ll 1$. For the first four diagrams, to sum up together, the imaginary part of the amplitude ${\cal A}(+,+,+)$ is \begin{equation} \label{im1} {\rm Im}{\cal A}^{1234}(+,+,+)=\alpha_k^2(1+\alpha_k){\cal N}\times \int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2) (\frac{2}{9}-\frac{-1}{36}\frac{k_T^2-(1+\alpha_k)(k_T,l_T)}{(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2}), \end{equation} where $\frac{2}{9}$ and $\frac{-1}{36}$ are the color factors for Diags.1,4 and Diags.2,3 respectively, and ${\cal N}$ is defined as \begin{equation} {\cal N}=\frac{3s}{\sqrt{-2\alpha_kk_T^2}}g_s^3T_{ij}^a. \end{equation} The other helicity amplitudes for the first four diagrams have the similar forms as (\ref{im1}), \begin{eqnarray} \label{im11} \nonumber {\rm Im}{\cal A}^{1234}(-,-,-)&=&{\rm Im}{\cal A}^{1234}(+,+,+),\\ {\rm Im}{\cal A}^{1234}(+,-,+)&=&{\rm Im}{\cal A}^{1234}(-,+,-)=\frac{-1}{\alpha_k}{\rm Im}{\cal A}^{1234}(+,+,+). \end{eqnarray} These amplitude expressions Eq.~(\ref{im1}) can also serve as the calculations of the amplitude for the diffractive direct photon production process $q p\rightarrow q\gamma p$ \cite{photon} except the difference on the color factors.\footnote{In Ref.\cite{photon}, we did not employ the helicity amplitude method. If we use the amplitude expressions Eqs.(\ref{im1}) and (\ref{im11}) (correct the color factors) to calculate the photon production process $qp\rightarrow q\gamma p$, we can get the same result as that in \cite{photon}. This can be viewed as a cross check for the methods we used in the calculations.} In the direct photon process, the color factors for these four diagrams are the same (they are all $\frac{2}{9}$). It is instructive to see what is the consequence of this difference. We know that the amplitude of the diffractive process in Eq.~(\ref{ima}) must be zero in the limit $l_T^2\rightarrow 0$. Otherwise, this will lead to a linear singularity when we perform the integration of the amplitude over $l_T^2$ due to existence of the factor $1/(l_T^2)^2$ in the integral of Eq.~(\ref{ima})\cite{charm}. This linear singularity is not proper in QCD calculations. So, we must first exam the amplitude behavior under the limit of $l_T^2\rightarrow 0$ for all the diffractive processes in the calculations using the two-gluon exchange model. From Eq.~(\ref{im1}), we can see that the amplitude for the diffractive direct photon production process $qp\rightarrow q\gamma p$ is exact zero at $l_T^2\rightarrow 0$. However, for the process $qp\rightarrow qgp$ the amplitude for the first four diagrams is not exact zero in the limit $l_T^2\rightarrow 0$ due to the inequality of the color factors between them. So, for this process there must be other diagrams in this order of perturbative QCD calculation to cancel out the linear singularity which rises from the first four diagrams. The last five diagrams of Fig.2 are just for this purpose. Finally, by adding up all of the nine diagrams of Fig.2, the imaginary parts of the amplitudes are \begin{eqnarray} \nonumber {\rm Im}{\cal A}(+,+,+)&=&{\rm Im}{\cal A}(-,-,-)=\frac{\alpha_k^2}{4}{\cal N}\times {\cal T},\\ {\rm Im}{\cal A}(+,-,+)&=&{\rm Im}{\cal A}(-,+,-)=-\frac{\alpha_k}{4}{\cal N}\times {\cal T}, \end{eqnarray} where \begin{eqnarray} \label{int2} \nonumber {\cal T}&=&\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)[ \frac{(1+\alpha_k)^2}{9}\frac{(k_T,l_T)-(1+\alpha_k)l_T^2}{(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2} -(1+\alpha_k)\frac{(k_T,l_T)+l_T^2}{(\vec{k}_T+\vec{l}_T)^2}\\ &&-\alpha_k\frac{(k_T,l_T)-l_T^2}{(\vec{k}_T-\vec{l}_T)^2} +\alpha_k^2\frac{(k_T,l_T)-\alpha_kl_T^2}{(\vec{k}_T-\alpha_k\vec{l}_T)^2}]. \end{eqnarray} From the above results, we can see that in the integration of the amplitude the linear singularity from different diagrams are canceled out by each other, which will guarantee there is no linear singularity in the total sum. Another feature of the above results for the amplitudes is the relation to the differential off-diagonal gluon distribution function $f(x',x'';l_T^2)$. However, as mentioned above that there is no big difference between the off-diagonal gluon distribution function and the usual gluon distribution at small $x$, so we can simplify the integration of (\ref{int2}) by approximating the differential off-diagonal gluon distribution function $f(x',x'';l_T^2)$ by the usual diagonal differential gluon distribution function $f_g(x;l_T^2)$. After integrating over the azimuth angle of $\vec{l}_T$, the integration ${\cal T}$ will then be \begin{eqnarray} \label{qt} \nonumber {\cal T}&=&\pi\int\frac{dl_T^2}{(l_T^2)^2}f_g(x;l_T^2)[ \frac{1+\alpha_k}{9}(\frac{1}{2}-\frac{k_T^2-(1+\alpha_k)l_T^2}{2|k_T^2-(1+\alpha_k)l_T^2|}) +(\frac{1}{2}-\frac{k_T^2-l_T^2}{2|k_T^2-l_T^2|})\\ &&+\alpha_k(\frac{1}{2}-\frac{k_T^2-\alpha_k l_T^2}{2|k_T^2-\alpha_k l_T^2|})]. \end{eqnarray} In the above integration, if $l_t^2<k_T^2/(1+\alpha_k)^2$ the first term of the integration over $l_T^2$ will be zero; if $l_t^2<k_T^2$ the second term will be zero; if $l_t^2<k_T^2/\alpha_k^2$ the third term will be zero. So, the dominant regions contributing to the three integration terms are $l_t^2\sim k_T^2/(1+\alpha_k)^2$, $l_t^2\sim k_T^2$, and $l_t^2\sim k_T^2/\alpha_k^2$ respectively. Approximately, by ignoring some evolution effects of the differential gluon distribution function $f_g(x;l_T^2)$ in the above dominant integration regions, we get the following results for the integration ${\cal T}$, \begin{equation} \label{i1} {\cal T}=\frac{\pi}{k_T^2}[f_g(x;k_T^2)+\frac{(1+\alpha_k)^3}{9}f_g(x;\frac{k_T^2}{(1+\alpha_k)^2}) +\alpha_k^3f_g(x;\frac{k_T^2}{\alpha_k^2})]. \end{equation} Obtained the formula for the integration ${\cal T}$, the amplitude squared for the partonic process $qp\rightarrow qgp$ will be reduced to, after averaging over the spin and color degrees of freedom, \begin{equation} \overline{|{\cal A}|}^2=\frac{\alpha_s^3(4\pi)^3}{24}\frac{1+\alpha_k^2}{M_X^2(1+\alpha_k)}s^2|{\cal T}|^2. \end{equation} And the cross section for the partonic process $qp\rightarrow qgp$ is \begin{eqnarray} \label{xsp} \nonumber \frac{d\hat\sigma(qp\rightarrow qgp)}{dt}|_{t=0}&=&\int_{M_X^4>4k_T^2}dM_X^2dk_T^2d\alpha_k[\delta(\alpha_k-\alpha_1)+\delta(\alpha_k-\alpha_2)]\\ &&\frac{\alpha_s^3}{96(M_X^2)^2}\frac{1+\alpha_K^2}{1+\alpha_k}\frac{1}{\sqrt{1-\frac{4k_T^2}{M_X^2}}} |{\cal T}|^2, \end{eqnarray} where $\alpha_{1,2}$ are the solutions of the following equations, \begin{equation} \alpha(1+\alpha)+\frac{k_T^2}{M_X^2}=0. \end{equation} The integral bound $M_X^2>4k_T^2$ in (\ref{xsp}) shows that the dominant contribution of the integration over $M_X^2$ comes from the region of $M_X^2\sim 4k_T^2$. Using Eq.~(\ref{ak}), this indicates that in this dominant region $\alpha_k$ is of order of 1. So, in the integration ${\cal T}$ the differential gluon distribution function $f_g(x;Q^2)$ of the three terms can approximately take their values at the same scale of $Q^2=k_T^2$. That is, the integration ${\cal T}$ is then simplified to \begin{equation} \label{i2} {\cal T}=\frac{\pi}{9k_T^2}f_g(x;k_T^2)(1+\alpha_k)(10-7\alpha_k+10\alpha_k^2). \end{equation} Numerical calculations show that there is little difference between the cross sections by using these two different parametrizations of ${\cal T}$, Eq.~(\ref{i1}) and Eq.~(\ref{i2}). So, in Sec.IV, we use Eqs.~(\ref{xsp}) and (\ref{i2}) to estimate the diffractive production rate at the Fermilab Tevatron. \subsection{$gp\rightarrow ggp$ process} For the partonic process $gp\rightarrow ggp$, there are twelve diagrams in the leading order contributions as shown in Fig.3. The first nine diagrams are due to the existence of the three-gluon interaction vertex, and the last three diagrams are due to the existence of the four-gluon interaction vertex. But it will be shown in the following calculations, the last three diagrams do not contribute under some choice of the polarizations of the three external gluons. The Sudakov variables can be calculated by the similar method used in the last subsection. And those Sudakov variables of the momenta $u$, $k$ for the process $gp\rightarrow ggp$ are the same as those in the last subsection. For the loop momentum $l$, the relevant Sudakov variables for each diagram are \begin{eqnarray} \nonumber \alpha_l&=&-\frac{l_T^2}{s},\\ \nonumber \beta_l&=&\frac{2(k_T,l_T)-l_T^2}{\alpha_ks},~~~{\rm for~Diag.}1,~4,~6,~10,\\ \nonumber &=&\frac{2(k_T,l_T)+l_T^2}{(1+\alpha_k)s},~~~{\rm for~Diag.}2,~3,~5,~11,\\ &=&-\frac{M_X^2-l_T^2}{s},~~~~~~~{\rm for~Diag.}7,~8,~9,~12. \end{eqnarray} And also, we can express the cross section formula for the partonic process $gp\rightarrow ggp$ in the following form, \begin{equation} \label{gxs} \frac{d\hat{\sigma}(gp\rightarrow ggp)}{dt}|_{t=0}=\frac{dM_X^2d^2k_Td\alpha_k}{16\pi s^216\pi^3M_X^2} \delta(\alpha_k(1+\alpha_k)+\frac{k_T^2}{M_X^2})\sum \overline{|{\cal A}|}^2, \end{equation} where ${\cal A}$ is the amplitude of the process $gp\rightarrow ggp$. We know that the real part of the amplitude ${\cal A}$ is zero, and the imaginary part of the amplitude ${\cal A}(gp\rightarrow ggp)$ for each diagram of Fig.3 has the following general form, \begin{equation} \label{gima} {\rm Im}{\cal A}=C_Ff_{abc}\int \frac{d^2l_T}{(l_T^2)^2}G(e_1,e_2,e_3)\times F, \end{equation} where $C_F$ is the color factor for each diagram. $a,~b,~c$ are the color indexes for the incident gluon and the two outgoing gluons respectively, and $f_{abc}$ are the antisymmetric $SU(3)$ structure constants. $G(e_1,e_2,e_3)$ represents the interaction part including one propagator for the first nine diagrams, where $e_1,~e_2,~e_3$ are the polarization vectors for the incident gluon and the two outgoing gluons. $F$ in the integral is the same as that in Eq.(\ref{feq}). The color factors $C_F$ for the twelve diagrams are \begin{eqnarray} \label{cf} \nonumber C_F&=&\frac{1}{2},~~~~~~~{\rm for~ Diag.}1,~7,\\ \nonumber C_F&=&-\frac{1}{2},~~~~~{\rm for~ Diag.}2,\\ \nonumber C_F&=&-\frac{1}{4},~~~~~{\rm for~ Diag.}3,~6,~9,\\ \nonumber C_F&=&\frac{1}{4},~~~~~~~{\rm for~ Diag.}4,~5,~8,\\ C_F&=&\frac{3}{4},~~~~~~~{\rm for~ Diag.}10,~11,~12. \end{eqnarray} Following the calculation method used in the last subsection, we employ the helicity amplitude method to calculate the amplitude Eq.(\ref{gima}). For the polarization vector of the incident gluon, which is transversely polarized, we choose, \begin{equation} \label{ev} e^{(\pm)}_1=\frac{1}{\sqrt{2}}(0,1,\pm i,0). \end{equation} For the two outgoing gluons, we choose their polarization vectors as\cite{wu} \begin{eqnarray} \label{geprc} \nonumber \not\! e_2^{(\pm)}=N_e[(\not\! k+\not\! q)\not\! q\not\! p(1\mp\gamma_5)+ \not\! p\not\! q(\not\! k+\not\! q)(1\pm\gamma_5)],\\ \not\! e_3^{(\pm)}=N_e[(\not\! u-\not\! k)\not\! q\not\! p(1\mp\gamma_5)+ \not\! p\not\! q(\not\! u-\not\! k)(1\pm\gamma_5)]. \end{eqnarray} The normalization factor $N_e$ has the same form as in Eq.(\ref{ene}). Under the above choice of the polarization vectors for the external gluons, we can easily find that they are satisfied the following equations, \begin{equation} p\cdot e_1=p\cdot e_2=p\cdot e_3=0. \end{equation} With these relations, we can further find that the last three diagrams do not contribute to the partonic process $gp\rightarrow ggp$. For the first nine diagrams, there are two helicity amplitudes among the eight helicity amplitudes do not contribute in the context of the above choice of the polarizations of the external gluons, i.e., \begin{equation} {\rm Im}{\cal A}(+,+,+)={\rm Im}{\cal A}(-,-,-)=0. \end{equation} In the expression of the amplitude ${\cal A}(\lambda(e_1),\lambda(e_2),\lambda(e_3))$, $\lambda$ denote the helicities for the three gluons respectively. The other six helicity amplitudes are divided into the following three different sets, \begin{eqnarray} \nonumber {\rm Im}{\cal A}(+,-,-)&\sim &{\rm Im}{\cal A}(-,+,+),\\ \nonumber {\rm Im}{\cal A}(+,-,+)&\sim &{\rm Im}{\cal A}(-,+,-),\\ {\rm Im}{\cal A}(+,+,-)&\sim &{\rm Im}{\cal A}(-,-,+). \end{eqnarray} For the first helicity amplitudes set, ${\rm Im}{\cal A}(\pm,\mp,\mp)$, to sum up all of the nine diagrams, we get \begin{equation} \label{gm1} {\rm Im}{\cal A}(\pm,\mp,\mp)={\cal N}'\pi \vec{e}_1^{(\pm)}\cdot k_T{\cal I}, \end{equation} where ${\cal N}'$ is defined as \begin{equation} {\cal N}'=\frac{3}{4}\frac{s}{k_T^2}g_s^3f_{abc}. \end{equation} And the integration ${\cal I}$ is \begin{eqnarray} \label{gim11} \nonumber {\cal I}&=&{1\over \pi}\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)[-(1+\alpha_k)\frac{k_T^2+(k_T,l_T)}{(\vec{k}_T+\vec{l}_T)^2} +\alpha_k\frac{k_T^2-(k_T,l_T)}{(\vec{k}_T-\vec{l}_T)^2} \\ &&+(1+\alpha_k)^2\frac{k_T^2-(1+\alpha_k)(k_T,l_T)}{(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2} -\alpha_k^2\frac{k_T^2-\alpha_k(k_T,l_T)}{(\vec{k}_T-\alpha_k\vec{l}_T)^2}]\\ \nonumber &=&{1\over \pi}\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)[(1+\alpha_k)\frac{(k_T,l_T)+l_T^2}{(\vec{k}_T+\vec{l}_T)^2} +\alpha_k\frac{(k_T,l_T)-l_T^2}{(\vec{k}_T-\vec{l}_T)^2} \\ &&-(1+\alpha_k)^2\frac{(k_T,l_T)-(1+\alpha_k)l_T^2}{(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2} +\alpha_k^2\frac{(k_T,l_T)-\alpha_kl_T^2}{(\vec{k}_T-\alpha_k\vec{l}_T)^2}]. \end{eqnarray} From the above equations, we can check that there is no linear singularity at the limit of $l_T^2\rightarrow 0$ in the integration of the amplitude over the loop momentum. The first term of the integration ${\cal I}$ in Eq.(\ref{gim11}) comes from the contribution of Diag.3; the second term comes from Diag.4; the third term comes from Diag.6 and Diag.9; the last term comes from Diag.5 and Diag.8. The contributions from Diags.1, 2 and 7 are canceled out by each other. From (\ref{gim11}), we can see that the linear singularities coming from the four terms are canceled out by each other. The final result for the amplitude is now free of linear singularity. We must emphasize here that only the total sum of the contributions from all of the diagrams is free of linear singularity. The separation of these diagrams will cause linear singularity. Following the argument in the last subsection of the calculation for the partonic process $qp\rightarrow qgp$, we can approximate the differential off-diagonal gluon distribution function $f(x',x'';l_T^2)$ by the usual diagonal differential gluon distribution function $f_g(x;l_T^2)$ to further simplify the integration of ${\cal I}$. After integrating over the azimuth angle of $\vec{l}_T$, this integration will then be \begin{eqnarray} \label{gi2} \nonumber {\cal I}&=&\int\frac{dl_T^2}{(l_T^2)^2}f_g(x;l_T^2)[(\frac{1}{2}-\frac{k_T^2-l_T^2}{2|k_T^2-l_T^2|}) -{(1+\alpha_k)}(\frac{1}{2}-\frac{k_T^2-(1+\alpha_k)l_T^2}{2|k_T^2-(1+\alpha_k)l_T^2|}) \\ &&+\alpha_k(\frac{1}{2}-\frac{k_T^2-\alpha_k l_T^2}{2|k_T^2-\alpha_k l_T^2|})]. \end{eqnarray} The above equation shows that the integration ${\cal I}$ here has the similar behavior as that of the integration ${\cal T}$ of Eq.(\ref{qt}) in the last subsection. So, the three terms of the above integration ${\cal I}$ are dominantly contributed from the integral regions of $l_T^2$ as $l_t^2\sim k_T^2/(1+\alpha_k)^2$, $l_t^2\sim k_T^2$, and $l_t^2\sim k_T^2/\alpha_k^2$ respectively. Approximately, we may also ignore the evolution effects of the differential gluon distribution function $f_g(x;l_T^2)$ in the above dominant integration regions, and so the integration ${\cal I}$ is reduced to \begin{equation} \label{gi1} {\cal I}=\frac{1}{k_T^2}[f_g(x;k_T^2)-(1+\alpha_k)^3f_g(x;\frac{k_T^2}{(1+\alpha_k)^2}) +\alpha_k^3f_g(x;\frac{k_T^2}{\alpha_k^2})]. \end{equation} For the second helicity amplitudes set, ${\rm Im}{\cal A}(\pm,\mp,\pm)$, the calculations are more complicated, and the contribution from Diag.3 is \begin{eqnarray} \label{g21} \nonumber {\rm Im}{\cal A}^3(\pm,\mp,\pm)&=&-{\cal N}'(1+\alpha_k)^2\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)\\ && \frac{\alpha_kk_T^2\vec{e}_1^{(\pm)}\cdot(\vec{k}_T+\vec{l}_T)+ (k_T^2+(k_T,l_T))\vec{e}_1^{(\pm)}\cdot\vec{k}_T}{(\vec{k}_T+\vec{l}_T)^2}. \end{eqnarray} The contribution from Diag.4 is \begin{eqnarray} \label{g22} \nonumber {\rm Im}{\cal A}^4(\pm,\mp,\pm)&=&{\cal N}'\alpha_k(1+\alpha_k)\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)\\ && \frac{\alpha_kk_T^2\vec{e}_1^{(\pm)}\cdot(\vec{k}_T-\vec{l}_T)+ (k_T^2-(k_T,l_T))\vec{e}_1^{(\pm)}\cdot\vec{k}_T}{(\vec{k}_T-\vec{l}_T)^2}. \end{eqnarray} The contributions from Diag.5 and Diag.8, to sum up together, are \begin{eqnarray} \label{g23} \nonumber {\rm Im}{\cal A}^{58}(\pm,\mp,\pm)&=&-{\cal N}'\alpha_k(1+\alpha_k)\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)\\ && \frac{\alpha_kk_T^2\vec{e}_1^{(\pm)}\cdot(\vec{k}_T-\alpha_k\vec{l}_T)+ (k_T^2-\alpha_k(k_T,l_T))\vec{e}_1^{(\pm)}\cdot\vec{k}_T}{(\vec{k}_T-\alpha_k\vec{l}_T)^2}. \end{eqnarray} The contributions from Diag.6 and Diag.9, to sum up together, are \begin{eqnarray} \label{g24} \nonumber {\rm Im}{\cal A}^{69}(\pm,\mp,\pm)&=&{\cal N}'(1+\alpha_k)^2\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)\\ && \frac{\alpha_kk_T^2\vec{e}_1^{(\pm)}\cdot(\vec{k}_T-(1+\alpha_k)\vec{l}_T)+ (k_T^2-(1+\alpha_k)(k_T,l_T))\vec{e}_1^{(\pm)}\cdot\vec{k}_T}{(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2}. \end{eqnarray} The contributions from other three diagrams (Diag.1, Diag.2 and Diag.7) are canceled out by each other. From the above results Eqs.(\ref{g21}-\ref{g24}), we can see that every term has linear singularity at the limit of $l_T^2\rightarrow 0$ in the integration of the amplitude over $l_T^2$, while their total sum is free of the linear singularity. Following the procedure as we do for the helicity amplitude ${\cal A}(\pm,\mp,\mp)$ in the above, we can approximate the off-diagonal gluon distribution function $f(x',x'';l_T^2)$ by the usual diagonal differential gluon distribution function $f_g(x;l_T^2)$. After integrating over the azimuth angle of $\vec{l}_T$, to sum up all of Eqs.(\ref{g21}-\ref{g24}), we get the helicity amplitude, \begin{equation} \label{gm2} {\rm Im}{\cal A}(\pm,\mp,\pm)={\cal N}'\pi (1+\alpha_k)^2\vec{e}_1^{(\pm)}\cdot k_T{\cal I}, \end{equation} where ${\cal I}$ is the same as Eq.(\ref{gi2}) and then Eq.(\ref{gi1}) under the same approximation. For the third helicity amplitudes set, ${\rm Im}{\cal A}(\pm,\pm,\mp)$, the calculations are similar to the calculations of ${\rm Im}{\cal A}(\pm,\mp,\pm)$. The contribution from Diag.3 is \begin{eqnarray} \label{g31} \nonumber {\rm Im}{\cal A}^3(\pm,\pm,\mp)&=&-{\cal N}'\alpha_k(1+\alpha_k)\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)\\ && \frac{(1+\alpha_k)k_T^2\vec{e}_1^{(\pm)}\cdot(\vec{k}_T+\vec{l}_T)- (k_T^2+(k_T,l_T))\vec{e}_1^{(\pm)}\cdot\vec{k}_T}{(\vec{k}_T+\vec{l}_T)^2}. \end{eqnarray} The contribution from Diag.4 is \begin{eqnarray} \label{g32} \nonumber {\rm Im}{\cal A}^4(\pm,\pm,\mp)&=&{\cal N}'\alpha_k^2\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)\\ && \frac{(1+\alpha_k)k_T^2\vec{e}_1^{(\pm)}\cdot(\vec{k}_T-\vec{l}_T)- (k_T^2-(k_T,l_T))\vec{e}_1^{(\pm)}\cdot\vec{k}_T}{(\vec{k}_T-\vec{l}_T)^2}. \end{eqnarray} The contributions from Diag.5 and Diag.8, to sum up together, are \begin{eqnarray} \label{g33} \nonumber {\rm Im}{\cal A}^{58}(\pm,\pm,\mp)&=&-{\cal N}'\alpha_k^2\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)\\ && \frac{(1+\alpha_k)k_T^2\vec{e}_1^{(\pm)}\cdot(\vec{k}_T-\alpha_k\vec{l}_T)- (k_T^2-\alpha_k(k_T,l_T))\vec{e}_1^{(\pm)}\cdot\vec{k}_T}{(\vec{k}_T-\alpha_k\vec{l}_T)^2}. \end{eqnarray} The contributions from Diag.6 and Diag.9, to sum up together, are \begin{eqnarray} \label{g34} \nonumber {\rm Im}{\cal A}^{69}(\pm,\pm,\mp)&=&{\cal N}'\alpha_k(1+\alpha_k)\int\frac{d^2\vec{l}_T}{(l_T^2)^2}f(x',x'';l_T^2)\\ && \frac{(1+\alpha_k)k_T^2\vec{e}_1^{(\pm)}\cdot(\vec{k}_T-(1+\alpha_k)\vec{l}_T)- (k_T^2-(1+\alpha_k)(k_T,l_T))\vec{e}_1^{(\pm)}\cdot\vec{k}_T}{(\vec{k}_T-(1+\alpha_k)\vec{l}_T)^2}. \end{eqnarray} And also, we find that the contributions from other three diagrams (Diag.1, Diag.2 and Diag.7) are canceled out by each other, and the total sum of Eqs.(\ref{g31}-\ref{g34}) is free of the linear singularity. If we approximate the off-diagonal gluon distribution function $f(x',x'';l_T^2)$ by the usual diagonal differential gluon distribution function $f_g(x;l_T^2)$, and integrate over the azimuth angle of $\vec{l}_T$, their sum will lead to a similar result as in Eq.(\ref{gm2}), \begin{equation} \label{gm3} {\rm Im}{\cal A}(\pm,\pm,\mp)={\cal N}'\pi \alpha_k^2\vec{e}_1^{(\pm)}\cdot k_T{\cal I}. \end{equation} By summing up all of the helicity amplitudes Eqs.(\ref{gm1}), (\ref{gm2}) and {\ref{gm3}), we will get the amplitude squared for the partonic process $gp\rightarrow ggp$, after averaging over the spin and color degrees of freedom, \begin{equation} \overline{|{\cal A}|}^2=\frac{27\pi^2\alpha_s^3(4\pi)^3}{16}\frac{s^2}{k_t^2}(1-\frac{k_T^2}{M_X^2})^2|{\cal I}|^2. \end{equation} And the cross section for the partonic process $gp\rightarrow ggp$ is \begin{equation} \label{gxsp} \frac{d\hat\sigma(gp\rightarrow ggp)}{dt}|_{t=0}=\int_{M_X^4>4k_T^2}dM_X^2dk_T^2 \frac{27\alpha_s^3\pi^2}{32M_X^2k_T^2}(1-\frac{k_T^2}{M_X^2})^2|{\cal I}|^2\frac{1}{\sqrt{1-\frac{4k_T^2}{M_X^2}}}, \end{equation} Following the same argument in the last subsection for the calculations of the partonic process $qp\rightarrow qgp$, we see that the dominant contribution of the integration over $M_X^2$ comes from the region of $M_X^2\sim 4k_T^2$, where the differential gluon distribution function $f_g(x;Q^2)$ of the three terms in the integration ${\cal I}$ can approximately take their values at the same scale of $Q^2=k_T^2$. That is, the integration ${\cal I}$ is then simplified to \begin{equation} \label{i22} {\cal I}=\frac{1}{k_T^2}(-3\alpha_k(1+\alpha_k))f_g(x;k_T^2) =\frac{1}{k_T^2}\frac{3k_T^2}{M_X^2}f_g(x;k_T^2)=\frac{3}{M_X^2}f_g(x;k_T^2). \end{equation} \section{Numerical results} In this section we study the numerical behavior of the diffractive gluon jet production at the Fermilab Tevatron. We will study the $p_T$ distribution and $x_1$ distribution of the cross section. We will also compare the gluon jet production with the quark jet production which has been calculated in \cite{charm,quark}. A more thorough phenomenological study, including a comparison to currently available data at Tevatron on the diffractive dijet production rate, will be presented elsewhere. Provided with the cross section formulas for the partonic processes $qp\rightarrow qg p$ (\ref{xsp}) and $gp\rightarrow ggp$ (\ref{gxsp}), we can calculate the cross section of the diffractive gluon jet production at hadron level. However, as mentioned above, there exists nonfactorization effect caused by the spectator interactions in the hard diffractive processes in hadron collisions. Here, we use a suppression factor ${\cal F}_S$ to describe this nonfactorization effect in the hard diffractive processes at hadron colliders\cite{soper,survive}. At the Tevatron, the value of ${\cal F}_S$ may be as small as ${\cal F}_S\approx 0.1$\cite{soper,tev}. That is to say, the total cross section of the diffractive processes at the Tevatron may be reduced down by an order of magnitude due to this nonfactorization effect. In the following numerical calculations, we adopt this suppression factor value to evaluate the diffractive production rate. In our calculations, the scales for the parton distribution functions and the running coupling constant are both set to be $Q^2=k_T^2$. For the parton distribution functions, we choose the GRV NLO set \cite{grv}. In Fig.4, we plot the differential cross section $d\sigma/dt|_{t=0}$ as a function of the lower bound of the transverse momentum of the gluon jet, $k_{T{\rm min}}$. This figure shows that the cross section is sensitive to the transverse momentum cut $k_{T{\rm min}}$. We plot separately the contributions from the two subprocesses, $qp\rightarrow qgp$ and $gp\rightarrow ggp$. By comparison, we also plot the cross section of the diffractive light quark jet production calculated in\cite{quark}. The three curves in this figure show that the contribution from the subprocess $gp\rightarrow ggp$ is two orders of magnitude larger than that from the subprocess $qp\rightarrow qgp$ for the diffractive gluon jet production, and the light quark jet production rate is in the same order with that of the subprocess $qp\rightarrow qgp$. This indicates that the diffractive dijet production at hadron colliders dominantly comes from the subprocess $gp\rightarrow ggp$ in the two-gluon exchange model. In Fig.5, we plot the differential cross section $d\sigma/dt|_{t=0}$ as a function of the lower bound of the momentum fraction of the proton carried by the incident gluon $x_{1{\rm min}}$, where we set $k_{T{\rm min}}=5~GeV$. Fig.5(a) is for the contribution from the subprocess $qp\rightarrow qgp$, and Fig.5(b) is from the subprocess $gp\rightarrow ggp$. These two figures show that the dominant contribution comes from the region of $x_1\sim 10^{-2}-10^{-1}$ for the subprocess $gp\rightarrow ggp$, and $x_1>10^{-1}$ for the subprocess $qp\rightarrow qgp$. These properties are similar to those of the diffractive charm jet and $W$ boson productions calculated in\cite{charm,dy}. \section{Conclusions} In this paper, we have calculated the diffractive gluon jet production at hadron colliders in perturbative QCD by using the two-gluon exchange model. We find that the production cross section is related to the squared of the differential gluon distribution function $\partial G(x;Q^2)/\partial ln Q^2$ at the scale of $Q^2\sim k_T^2$, where $k_T$ is the transverse momentum of the final state gluon jet. We have also compared the production rate of the gluon jet in the diffractive processes with those of the light quark jet and heavy quark jet productions, and found that the production rates of these processes are in the same order of magnitude. As we know, the large transverse momentum dijet production in the diffractive processes at hadron colliders is important to study the diffractive mechanism and the nature of the Pomeron. The CDF collaboration at the Fermilab Tevatron have reported some results on this process\cite{tev}. Up to now, we have calculated all of the dijet production subprocesses in the diffractive processes at hadron colliders, including $gp\rightarrow q\bar qp$, $qp\rightarrow qgp$ and $gp\rightarrow ggp$ processes. In a forthcoming paper, we will compare the available data on the diffractive dijet production cross section at the Tevatron\cite{tev} to the predictions of our model to test the validity of perturbative QCD description of the diffractive processes at hadron colliders. \acknowledgments This work was supported in part by the National Natural Science Foundation of China, the State Education Commission of China, and the State Commission of Science and Technology of China.
\section{Present observational constraints} The bulk of the mass of the {\sc lmc}\xspace resides in a nearly face-on disk, with an inclination usually taken to equal the canonical value of $i=33^{\circ}$ (\cite{westerlund}), although both lower ($27^{\circ}$) and higher (up to $45^{\circ}$) values have also been derived from morphological or kinematical studies of the {\sc lmc}\xspace. This disk is observed to rotate with a circular velocity $V_{C} \sim 80$~km/s out to at least $8^{\circ}$ from the {\sc lmc}\xspace center (\cite{schommer}). If all the stars belong to the same population, with a vertical ({\it i.e.} perpendicular to the disk) velocity dispersion $\sigma_{W}$, the microlensing optical depth of such a disk upon its own stars is given by $\tau \sim 2 \sigma_{W}^{2} \sec^{2}i / c^{2} $ (\cite{gou95}). Considering the measured velocity of {\sc lmc}\xspace carbon stars (Cowley \& Hartwick 1991), Gould (1995) assumed $\sigma_{W} = 20$ km/s as a typical velocity dispersion for {\sc lmc}\xspace stars. He thus concluded that $\tau \sim 10^{-8}$, {\it i.e.} that self-lensing (first suggested by \cite{sahu94} and \cite{wu94}) contributes very little to the observed optical depth towards this line of sight. Carbon stars however may not be the ultimate probe to infer the velocity dispersion of {\sc lmc}\xspace populations: they actually comprise various ill-defined classes of objects (Me\-nes\-sier 1999), and their prevalence is a complex function of age, metallicity and probably other factors (Gould 1999). Both observational and theoretical arguments favour the existence of a wide range of velocity dispersions among the various {\sc lmc}\xspace stellar populations. To commence, Mea\-theringham {et al.}\xspace (1988) have determined the radial velocities of a sample of planetary nebulae (PN) in the {\sc lmc}\xspace. They measured a velocity dispersion of 19.1 km/s, much larger than the value of 5.4 km/s found for the HI. This was interpreted as being suggestive of orbital heating and diffusion operating in the {\sc lmc}\xspace in the same way as it is observed in the solar neighbourhood. Then, the observations of Hughes {et al.}\xspace (1991) show clear evidence for an increase in the velocity dispersion of long period variables (LPV) as a function of their age. For young LPVs, the velocity dispersion is 12 km/s whereas for old LPVs, it reaches 35 km/s. More recently, Zaritsky {et al.}\xspace (1999) found a velocity dispersion of $\sigma = 18.4 \pm 1.4$ km/s for 190 vertical red clump (VRC) stars\footnote{see Zaritsky {et al.}\xspace (1999) and Beaulieu and Sackett (1998) for a definition of RC and VRC stars.} whereas for the red clump (RC), they measured a value of $\sigma = 32.2 \pm 3.8$ km/s on a sample of 75 objects (throughout this paper, error bars are converted from Zaritsky's 95 \% confidence levels to standard $1 \sigma$). A general trend appears: the velocity dispersion is an increasing function of the age. Just like for our own Milky Way, stars of the {\sc lmc}\xspace disk have been continuously undergoing dynamical scattering by, for instance, molecular clouds or other gravitational inhomogeneities. This results in an increase of the velocity dispersion of a given stellar population with its age, as will be further discussed in section 3. Notice that the main argument in disfavour of a {\sc lmc}\xspace self-lensing explanation is precisely the low value of the measured vertical velocity dispersions. However, the stellar populations so far surveyed predominantly consist of red giants. They are shown in the next section not to be representative of the bulk of the {\sc lmc}\xspace disk stars, and actually biased towards young ages: they are on average $\sim$ 2 Gyr old, to be compared to an {\sc lmc}\xspace age of $\sim$ 12 Gyr. \section{The age bias} The red clump population will illustrate the main thrust of our argument. Clump stars have burning helium cores whose size is approximately independent of the total mass of the object. They also have the same luminosity and hence they spend a fixed amount of time $\tau_{\rm \, He}$ in the clump, irrespective of their mass $m$. Such objects are evolved post-MS stars, which does not mean that they are necessarily old. We have assumed a Salpeter Initial Mass Function for the various {\sc lmc}\xspace stellar populations \begin{equation} {\displaystyle \frac{dN}{dm}} \propto m^{\displaystyle - \left( 1 + \alpha \right)} \;\; , \end{equation} with $\alpha = 1.35$. The stellar formation history has been borrowed from Geha {et al.}\xspace (1998). Their preferred model (e) corresponds to a stellar formation rate ${\cal F}(t)$ that has remained constant for 10 Gyr since the formation of the {\sc lmc}\xspace 12 Gyr ago. Then, two Gyr ago, ${\cal F}(t)$ has increased by a factor of three. The number of stars that formed at time $t$ and whose mass is comprised between $m$ and $m + dm$ may be expressed as \begin{equation} {\displaystyle \frac{d^{2} N}{dm \, dt}} = {\cal F}(t) \, m^{\displaystyle - \left( 1 + \alpha \right)} \;\; . \end{equation} We have assumed a mass-luminosity relation $L \propto m^{\beta}$ on the MS so that the stellar lifetime may be expressed as $\tau_{\rm MS} (m) = {12 \; {\rm Gyr}} / {m^{\beta - 1}}$ (since $\tau \propto m/L$). With these oversimplified but natural assumptions, a star whose initial mass is $\leq 1 \, {\rm M}_\odot$ is still today on the MS and cannot have reached the clump. Conversely, a heavier star with $m \geq 1 \, {\rm M}_\odot$ may well be today in a helium core burning stage provided that its formation epoch lies in the range between $t = - \, \tau_{\rm MS} (m)$ (the object has just begun core helium burning) and $t = - \, \tau_{\rm MS} (m) - \, \tau_{\rm \, He} (m)$ (the star is about to leave the red clump). The number of RC stars observed today with progenitor mass in the range between $m$ and $m + dm$ is therefore given by \begin{equation} dN_{\rm RC} = {\cal F} ( - \tau_{\rm MS} (m) ) \times m^{\displaystyle - \left( 1 + \alpha \right)} \, dm \times \tau_{\rm \, He} \;\; . \label{CHSTAR_1} \end{equation} To get more insight into the age bias at stake, we can parameterize the progenitor mass $m$ in terms of the age $\tau \equiv \tau_{\rm MS} (m)$. The previous relation simplifies into \begin{equation} \frac{dN_{\rm RC}}{d\tau} \; = \; {\displaystyle \frac{{\cal F} ( - \tau ) \, \tau_{\rm \, He}}{(\beta - 1)}} \; \tau^{\displaystyle \left( \gamma - 1 \right)} \;\; , \end{equation} where $\gamma = \alpha / (\beta - 1)$. This may be directly compared to the age distribution of the bulk of the {\sc lmc}\xspace stars that goes like ${\cal F} ( - \tau )$. With a Salpeter mass function and $\beta = 4.5$, we get a value of $\gamma = 0.4$. The excess of young RC stars goes as $1 / \tau^{0.6}$ and the bias is obvious. Other IMF are possible and a spectral index as large as $\alpha \sim \beta - 1 \sim 3.5$ would be required to invalidate the effect. HST data analyzed by Holtzman {et al.}\xspace (1997) nevertheless point towards a spectral index $\alpha$ that extends from 0.6 up to 2.1 for stars in the mass range $0.6 \leq m \leq 3$ ${\rm M}_\odot$. The average value corresponds actually to a Salpeter law. There has been furthermore a recent burst in the {\sc lmc}\xspace stellar formation rate. In order to model it, we may express the total number of today's RC stars as an integral where the progenitor mass $m$ runs from $m_{1} = 1 \, {\rm M}_\odot$ up to the tip of the IMF whose actual value is irrelevant and has been set equal to infinity here for simplicity. Notice that the specific progenitor mass $m_{2} \simeq 1.7 \, {\rm M}_\odot$ corresponds to stars born 2 Gyr ago, when the stellar formation rate increased by a factor of 3. Stars which formed before that epoch will be referred to as old. Their number is given by \begin{equation} N_{\rm RC}^{\rm old} = {\displaystyle \int_{m_{1}}^{m_{2}}} \, {\cal F} \left( - \tau_{\rm MS} \right) \, m^{\displaystyle - \left( 1 + \alpha \right)} \, dm \, \tau_{\rm \, He} \;\; . \end{equation} On the other hand, the number $N_{\rm RC}^{\rm young}$ of young clump stars is obtained similarly, with masses in excess of $m_{2}$. We readily infer a fraction of young stars \begin{equation} {N_{\rm RC}^{\rm young}} / N_{\rm RC} = {\displaystyle \frac{3}{2 \, + \, (m_{2}/m_{1})^{\alpha}}} \simeq 0.751 \;\; . \end{equation} Three quarters of the clump stars observed today in the {\sc lmc}\xspace have thus formed less than 2 Gyr ago, during the recent period of stellar formation mentioned above. Integrating $\tau_{\rm MS}$ over the RC population \begin{equation} \left< \tau \right> = \frac{1}{N_{\rm RC}} \; {\displaystyle \int_{m_{1}}^{\infty}} \tau_{\rm MS} \, dN_{\rm RC} \;\; , \end{equation} yields the average age \begin{equation} \left< \tau \right> = \left( 12 \; {\rm Gyr} \right) \times \frac{\alpha}{\alpha + \beta - 1} \times {\displaystyle \frac {m_{1}^{1 - \alpha - \beta} \, + 2 \, m_{2}^{1 - \alpha - \beta}} {m_{1}^{- \alpha} \, + 2 \, m_{2}^{- \alpha}}} \; . \end{equation} This gives a numerical value of $\sim 1.95$ Gyr. We thus conclude that today's clump stars are, on average, much younger than the {\sc lmc}\xspace disk. \section{Distributions of velocity dispersions} This simple analytical result has been checked by means of a Monte Carlo study. We have randomly generated a sample of $10^{8}$ {\sc lmc}\xspace stars. The progenitor mass was drawn in the range $0.1 \leq m \leq 10$ ${\rm M}_\odot$ according to a Salpeter law. The age of formation was drawn in the range $-12$~Gyr $\leq t \leq 0$ according to the stellar formation history ${\cal F}(t)$ favoured by Geha {et al.}\xspace (1998). The vertical velocity dispersion $\sigma_{W}$ was then evolved in time from formation up to now according to Wielen's (1977) relation: \begin{equation} \sigma_{W}^{2} \; = \; \sigma_{0}^{2} + C_{W} \, t. \label{diffeq} \end{equation} This purely diffusive relation is known to be inadequate to describe velocity dispersions in our Galaxy (Edvardsson et al. 1993). We will however use it in our model, as heating processes in the {\sc lmc}\xspace may be different than those in the galaxy. The {\sc lmc}\xspace is indeed subject to tidal heating by the Milky Way (\cite{weinberg99}) and has most probably suffered encounters with the {\sc smc}\xspace . Although this simple relation lacks a theoretical motivation, it will be shown to account for several features of the velocity distributions in the {\sc lmc}\xspace, without being at variance with any observation. The initial velocity dispersion $\sigma_{0}$ was taken to be 10~km/s, and the diffusion coefficient in velocity space along the vertical direction $C_{W}$ to be 300~$\unit{km}^{2} \unit{s}^{-2} \unit{Gy}^{-1}$ so that our oldest stars have a vertical velocity dispersion reaching up to $\sigma_{W}^{\rm MAX} = 60$ km/s. For each star, the actual vertical velocity was then randomly drawn, assuming a Gaussian distribution with width $\sigma_{W}$. In order to compare our Monte Carlo results with the Zaritsky {et al.}\xspace (1999) measurements of the radial velocities of {\sc lmc}\xspace clump stars, we selected two groups of stars according to their position in the HR diagram. Following Zaritsky {et al.}\xspace, we use their colour index \begin{equation} C \; \equiv \; 0.565 \, (B - I) \; + \; 0.825 \, (U - V + 1.15) \;\; , \end{equation} so that the RC population is defined by $3.1 < C < 3.4$ with a magnitude $19 < V < 19.3$ whereas the VRC stars have the same colour index C and brighter magnitudes $18 < V < 18.75$. In order to infer the colours and magnitudes of the stars that we generated, we used the isochrones computed by Bertelli {et al.}\xspace (1994) for a typical {\sc lmc}\xspace metallicity and helium abundance of $Z = 0.008$ and $Y = 0.25$. \begin{figure}[h!] \begin{center} \epsfig{file=Bd291.f1.eps,width=7.8cm} \caption{ Velocity distribution for a sample of 190 vertical red clump stars that have been generated by the Monte Carlo discussed in the text. That histogram is similar to Fig.~10 of Zaritsky {et al.}\xspace (1999). A velocity dispersion of 18~km/s is found for the full sample (solid smooth curve). } \label{fig_vrc} \end{center} \end{figure} A random sample of 190 stars that passed the VRC selection criteria is presented in Fig.~\ref{fig_vrc} where the vertical velocities are displayed. This histogram may be compared to Fig.~10 of Zaritsky {et al.}\xspace (1999) where no VRC star is found with a velocity in excess of 60 km/s. With the full statistics, our Monte Carlo generated a population of $\sim$ 2,900 VRC objects whose vertical velocity distribution has a RMS of $\sim$ 18 km/s. The agreement between the Zaritsky {et al.}\xspace observations and our Monte Carlo results is noteworthy. The average age of our VRC sample is $\sim$~0.87~Gyr. \begin{figure}[h!] \begin{center} \epsfig{file=Bd291.f2.eps,width=7.8cm} \caption{ Like in the previous figure, a distribution of 75 red clump stars is now featured. We inferred a velocity dispersion of 23 km/s for the full sample (solid smooth curve). Our distribution is similar to that presented in Fig.~11 of Zaritsky {et al.}\xspace (1999). No star exhibits a velocity larger than 70~km/s. } \label{fig_rc} \end{center} \end{figure} We also selected a random sample of 75 RC stars whose velocity distribution is featured in Fig.~\ref{fig_rc}. Even with a diffusion coefficient as large as $C_{W} = 300 \unit{km}^{2} \unit{s}^{-2} \unit{Gy}^{-1}$ so as to comply with a large {\sc lmc}\xspace self-lensing optical depth, our full statistics of 18,000 RC objects has a velocity dispersion of $\sim$ 23 km/s. This is slightly below the value of $\sigma = 32.2 \pm 3.8$ km/s quoted by Zaritsky {et al.}\xspace Observations are nevertheless fairly scarce with only 75 RC stars. When Zaristsky {et al.}\xspace fitted a Gaussian to the RC radial velocity distribution featured in the Fig.~11 of their paper, they obtained a 95~\% C.L. dispersion of $\sigma = 32^{+19}_{-16}$ km/s with a large uncertainty. Our Monte Carlo velocity dispersion of 23 km/s is definitely compatible with that result. We infer an average age for the RC population of $\sim$~1.8 Gyr to be compared to our analytical result of $\sim$ 1.95 Gyr. This agrees well with Beaulieu and Sackett's conclusion that isochrones younger than 2.5 Gyr are necessary to fit the red clump. Notice finally that our age estimates for these various clump populations are in no way related to {\sc lmc}\xspace kinematics. They merely result from the postulated Salpeter IMF, the Geha {et al.}\xspace preferred stellar formation history and the Bertelli {et al.}\xspace isochrones. With this model, 70\% in mass of the {\sc lmc}\xspace disk consists of objects whose vertical velocity dispersion is in excess of 25 km/s, although the average vertical velocity dispersion of RC stars, for instance, is only $\sim$ 23 km/s. What about the other measurements? The velocity dispersion of PNs has been found equal to 19.1 km/s (Meatheringham {et al.}\xspace 1988). These authors estimate that the bulk of the PNs have an age near 3.5 Gyr. They also note that younger objects are present down to an age of order $0.5 - 1.3$ Gyr. Meatheringham {et al.}\xspace come finally to the conclusion that the indicative age of the PN population is 2.1 Gyr. This value agrees well once again with our analytical estimate. Our Monte Carlo gives a slightly larger value of 2.4 Gyr for the age of the PNs, with a velocity dispersion of 24.7 km/s. Because the observed sample contains 94 objects, the measured value of 19.1 km/s suffers presumably from significant uncertainties. Quite interesting also are the measurements by Hughes {et al.}\xspace (1991) of the velocity dispersions of LPVs as a function of their age. Their sample of 63 ``old'' LPVs has a velocity dispersion of $\sigma = 35^{+10}_{-4}$ km/s. For the bulk of the {\sc lmc}\xspace populations, we obtain an average velocity dispersion of $\sim 37$ km/s. The problem at stake is actually the age of those old LPVs. These stars indeed display an age-period relation. However, Hughes {et al.}\xspace derived this relation from kinematics considerations, using precisely Eq. \ref{diffeq}, and postulating the same diffusion coefficient as in the Milky Way. They thus inferred an average age of 9.5~Gyr. Finding instead the position of these stars in a colour-magnitude diagram and using {\sc lmc}\xspace isochrones would have led to a clean determination of the age-period relation. A direct determination of the age of LPVs is nevertheless spoilt by a few biases. Some LPVs are carbon stars and the ejected material around them may considerably dim their luminosities. These stars may also pulsate on an harmonic of the fundamental mode. Both effects lead to an under-determination of their luminosity and hence to an overestimate of their age (Menessier 1999). As a matter of fact, Groenewegen and de Jong (1994) conclude that {\sc lmc}\xspace stars whose progenitor mass is less than 1.15 ${\rm M}_\odot$ never reach the instability strip on the AGB. This yields an upper limit on the age of LPVs of $\sim 7.3$ Gyr, in clear contradiction with the average age of 9.5 Gyr inferred by Hughes {et al.}\xspace for old LPVs. Finally, Schommer {et al.}\xspace (1992) have obtained a velocity dispersion of $21 - 24$ km/s for 9 old {\sc lmc}\xspace clusters. Their large $1\sigma$ error of $\sim$ 10 km/s is due to the small size of the sample. It is not clear whether or not these clusters have formed in the disk. If they nevertheless had, they would have undergone a fairly restricted orbital heating with respect to the {\sc lmc}\xspace stars. Those systems and the giant molecular clouds have actually comparable masses and the energy exchange between them does not result in a significant increase of the velocity dispersion of the clusters unlike what happens to the stars. \section{Multi-component model of the {\sc lmc}\xspace} We model the {\sc lmc}\xspace to contain several stellar populations, each associated with a different velocity dispersion $\sigma_{W,i}$ which has evolved according to Eq. \ref{diffeq}. We describe each of the ten components of our model by an ellipsoidal density profile \begin{equation} \rho_i(R,z) = \frac{\Lambda_i}{R^2 \; + \; {\displaystyle {z^2}/{(1-e_i^2)}}} \;\; , \end{equation} up to a cut-off radius $R_{\rm MAX} = 15 \unit{kpc}$ (Aubourg et al. 1999). The multi-component model based on these profiles is self-consistent in the sense that it satisfies Poisson equation and results in a flat rotation curve with the desired $V_C$ of 80~km/s. We define the set of $\sigma_{W,i}$ so as to sample linearly the range between $\sigma_0 = 10 \unit{km/s}$ and $\sigma_{W}^{\rm MAX} = 60 \unit{km/s}$ (see previous section). The parameters $\Lambda_i$ and the ellipticities $e_i$ are determined so that the model reproduces the set of velocity dispersions $\sigma_{W,i}$ and surface mass densities $\Sigma_i$ where $d\Sigma_i/d\sigma_i \propto \sigma_i {\cal F}(t)$ with ${\cal F}(t)$ the stellar formation history of the {\sc lmc}\xspace mentioned in section 2. Assuming a typical M/L of 3, which is a free parameter in our model, we reproduce the observed surface brightness of the LMC. For a given distribution of objects, one can compute the total self-lensing optical depth $\tau$ and the event rate $\Gamma$. Both quantities are integrated on all deflectors and sources, considering that only main sequence stars brighter than $V = 20$ and red giants can be potential sources, since they are the only objects bright enough to be visible in microlensing surveys. The computation of $\Gamma$ requires an estimate of the relative transverse velocity of deflector and source, for which we have assumed an horizontal velocity dispersion equal to the vertical one predicted by the model. Details of this computation can be found in (\cite{nous}). For the model described above, one obtains $\tau = 9.3 \times 10^{-8}$ and $\Gamma = 3.5 \times 10^{-7} \unit{yr}^{-1}$. This can be compared to the {\sc eros}\xspace and {\sc macho}\xspace optical depths, respectively $8.2 \times 10^{-8}$ (\cite{eros1}) and $2.9_{-0.9}^{+1.4} \times 10^{-7}$ (Alcock {et al.}\xspace 1997). A combination of those two results yields an average optical depth of $2.1_{-0.8}^{+1.3} \times 10^{-7}$ (\cite{BennettReport}), but preliminary {\sc macho}\xspace results from their five-year analysis (Sutherland 1999) hint to a reduced optical depth as compared to their two-year analysis. The model prediction is thus in good agreement with the results obtained so far from microlensing experiments. Another relevant prediction of the model is the distribution of event durations, $d\Gamma/d\Delta t$. Figure~\ref{dgamdtau} illustrates this prediction for our model, along with the distribution of observed {\sc macho}\xspace events. \begin{figure}[h] \begin{center} \epsfig{file=Bd291.f3.eps,width=7.8cm} \caption{ Predicted distribution of event durations $d\Gamma/d\Delta t$, superimposed with the {\sc macho}\xspace experimental distribution. The events are those presented by {\sc macho}\xspace at the IVth Microlensing Workshop (\cite{machoEvt}), corrected for blending and efficiency using the formulae in \cite{Macho2yr}. } \label{dgamdtau} \end{center} \end{figure} Our model thus reproduces both the total observed optical depth towards the {\sc lmc}\xspace and the observed event duration distribution, while complying with the velocity dispersion measurements. A self-lensing interpretation of {\em all} the microlensing events observed so far towards the {\sc lmc}\xspace thus appears to be a plausible explanation. \begin{acknowledgements} We wish to thank M.O. Menessier for useful discussions, and the members of the EROS collaboration for their comments. We thank Andy Gould, our referee, for his useful remarks and suggestions. \end{acknowledgements}
\section{Introduction} \label{sec:1} Light-front frame is believed to be a useful tool for solving bound state problems in QCD. Generally, bound state calculations in the field theory have two sources of complexity - they are relativistic and of many-body type. The method of flow equations copes with both of them, at least to the definite order in perturbation theory. One transforms the Hamiltonian to eliminate interactions changing particle number, reducing thus the bound state problem to a few-body problem. Simultaneously utraviolet divergencies occur, originating from the high-energy region. To complete renormalization one uses either coupling coherence or fixes counterterms to provide finite values for physical observables and to retain symmetries violated by the procedure~\cite{Perry}. Both type of flow equations of renormalization type and for the new (particle number conserving) interactions appear together. This program can be fulfilled in perturbation theory expansion. As a result, the bound state problem is approximated by a set of renormalized, effective interactions that do not change particle number. The sensitive tool in the light-front frame to check how accurately one desribes bound states by these effective interactions is to measure the violation of rotational symmetry. This symmetry is linked to a dynamical operator on the light-front, since rotations are dynamical, i.e. depend on the interaction. The symmetry may be spoiled in two steps: first, regularization and renormalization; and second, reduction to the effective few-body interactions with particle number conservation. The nonperturbative renormalization flow is of crucial importance for QCD~\cite{BrPe,Perry,pau96}, but one can disregard this point in QED bound state calculations, that disentangles the two problems mentioned above. To the leading order the results for QED are obtained in~\cite{JoPeGl}, where positronium system is described approximately by the effective electron-positron interaction. In the nonrelativistic limit the results for positronium spectrum agree with the results of covariant calculations. There are at least two other alternative approaches to solve for bound states in the light-front dynamics, the scheme of similarity renormalization of Glazek and Wilson~\cite{GlWi} and the method of iterated resolvents of Pauli~\cite{pau96}. In both schemes calculations of the effective electron-positron interaction are performed and the question of rotational invariance for positronium spectrum is investigated. Calculations done so far in the similarity renormalization scheme use the nonrelativistic limit to find corresponding eigenvalues in the bound state perturbation theory~\cite{BrPe2}. Analytical calculations are performed there for ground state: ground triplet levels are degenerate, indicating that rotational symmetry is restored~\cite{BrPe2}. Performing similarity renormalization one eliminates high-energy modes and absorbs relativistic effects into an effective band-diagonal Hamiltonian, which describes bound state creation at nonrelativistic energy scales. It is a well working scheme for such systems as positronium~\cite{Perry}; therefore nonrelativistic approximations done in this approach to extract eigenvalues from effective Hamiltonian are quite natural there. In general, it is not always the case. In fact rotational symmetry becomes kinematic one, like light-front boost in the nonrelativistic limit, i.e. total momentum and its projection can be considered approximately as quantum numbers, that makes simpler to trace rotational invariance in these calculations. In the method of iterated resolvents an effective electron-positron potential is obtained and exact numerical solution of positronium bound state equation with the given potential is done~\cite{TrPa,TrPa2}. Degenerate multiplets for ground as well as for exited states are obtained~\cite{TrPa}. \footnote{ Numerical solution of positronium bound state problem in the light-front frame can be found also in ~\cite{kpw92,KaPi}. } It is convenient to perform relativistic calculations in the light-front frame, which effectively has nonrelativistic kinematics. In the present work we perform relativistic few-body calculations for positronium spectrum numerically in the spirit of the work Trittmann et.al.~\cite{TrPa} (and using numerical code ~\cite{TrPa2}), based on the effective electron-positron Hamiltonian obtained by the flow equations~\cite{previous}. Effective interaction was derived there for different cutoff functions. The requirement of block-diagonalization of the Hamiltonian determines the generator only up to a unitary transformation of the blocks; this explains why the effective interaction may depend on the cutoff function. The question we investigate is to what extent rotational invariance is violated on the level of positronium spectrum and how does it depend on the choice of the cutoff function. We are not able to trace rotational invariance during the calculations, since it is dynamical operator. This is an excellent test for the method of flow equations itself and the control of approximations done during the calculations. \section{Formulation of the problem} \label{sec:2} We address to solve a light-front Hamiltonian bound state equation \begin{equation} H \vert\psi\rangle =E \vert\psi\rangle \label{eq:i1}\end{equation} for positronium. Using flow equations we transform the QED Hamiltonian $H$ to a block-diagonal effective Hamiltonian, which reduces positronium problem to a bound state problem in the electron-positron sector. The effective Hamiltonian for an electron and a positron is \begin{equation} H_{\rm eff}=H_0+U_{\rm eff} \label{eq:i2}\end{equation} where $H_0$ is the kinetic energy, and $U_{\rm eff}$ includes effective interactions generated by the flow equations in the second order in coupling constant. The integral bound state equation is written \begin{eqnarray} E \langle p_1,p_2; \lambda_1,\lambda_2 \vert \psi \rangle &=& (E_{p_1}\!+\!E_{p_2}) \langle p_1,p_2; \lambda_1,\lambda_2 \vert \psi \rangle \nonumber\\ &+& \sum_{\lambda_1^\prime,\lambda_2^\prime} \!\int\! d^3p^\prime_1 d^3p^\prime_2 \langle p_1,p_2; \lambda_1,\lambda_2 \vert U_{\rm eff}\vert p^\prime_1,p^\prime_2; \lambda_1^\prime,\lambda_2^\prime \rangle \langle p^\prime_1,p^\prime_2;\lambda_1^\prime, \lambda_2^\prime \vert \psi \rangle \nonumber\\ \label{eq:i3}\end{eqnarray} where the effective Hamiltonian pickes out from the positronium wave function $\vert \psi \rangle$ the lowest $e\bar{e}$-component $\langle p_1,p_2;\lambda_1,\lambda_2 \vert \psi \rangle$ with $p_i,\lambda_i$ being the light-front three-momenta and helicities, respectively, carried by an electron ($i=1$), and a positron ($i=2$). The primed quantities refer to the initial state, the unprimed ones to the final state. The effective interaction $ \langle p_1,p_2;\lambda_1,\lambda_2 \vert U_{\rm eff} \vert p^\prime_1,p^\prime_2;\lambda_1^\prime,\lambda_2^\prime \rangle = U_{\rm eff} \delta(p_1+p_2-p'_1-p'_2) $ will be specified below. In order to deduce a Lorentz invariant energy we consider the bound state equation written for operator $P^-P^+$, corresponding to the invariant mass-squared $M^2$ on the light-front, rather than for the light-front Hamiltonian operator $H=P^-$. The light-front integral equation, \\ \begin{eqnarray} M^2\ \langle x,\vec \kappa_{\!\perp}; \lambda_1, \lambda_2 \vert \psi \rangle &=& {m^2 \!+\! \vec \kappa_{\!\perp}^2 \over x(1\!-\!x)} \ \langle x,\vec \kappa_{\!\perp}; \lambda_1, \lambda_2 \vert \psi \rangle \nonumber\\ &+& \sum _{ \lambda_1^\prime,\lambda_2^\prime} \int\limits_D\!dx^\prime d^2 \vec \kappa_{\!\perp}^\prime\, \,\langle x,\vec \kappa_{\!\perp}; \lambda_1, \lambda_2 \vert V_{\rm eff} \vert x^\prime,\vec \kappa_{\!\perp}^\prime; \lambda_1^\prime, \lambda_2^\prime\rangle \ \langle x^\prime,\vec \kappa_{\!\perp}^\prime; \lambda_1^\prime,\lambda_2^\prime \vert \psi \rangle \nonumber\\ \label{eq:r96} \end {eqnarray} is independent of the total momentum $P^+$ and $\vec P_\perp$. We introduced $V_{\rm eff}=P^{+2}U_{\rm eff}$. In that equation only intrinsic transversal momenta $\vec \kappa_{\!\perp}$ and longitudinal momentum fractions $x = p_1^+/P^+$ appear ($p_1^\mu=(xP^+,x\vec P_\perp + \vec \kappa_{\!\perp},p_1^-)$). Its spectrum is thus manifestly independent of the kinematical state of the bound system, particularly of $P^+$ and $\vec{P}_{\perp}$, which reflects on the boost invariance peculiar to the light-front form~\cite{bpp97}. The integration domain $D$ is restricted by the covariant cutoff condition of Brodsky and Lepage~\cite{LeBr}, \begin{eqnarray} \frac{m^2+\vec{\kappa}_{\perp}^{2}}{x(1-x)}\leq \Lambda^2+4m^2 \,,\label{eq:domain}\end{eqnarray} which allows for states having a kinetic energy below the bare cutoff~$\Lambda$. The effective interaction between electron and positron, being a kernel in the integral equation~(\ref{eq:r96}), is generated by the flow equations~\cite{previous} \begin{eqnarray} V_{\rm eff} = &-& \frac{\alpha}{4\pi^2} \langle \gamma^\mu\gamma^\nu\rangle_{ex} \left[g_{\mu\nu}\, \left(\frac{\Theta_{e \bar e}} {Q_{e}^2} + \frac{\Theta_{\bar e e}} {Q_{\bar e}^2}\right) + \eta_\mu\eta_\nu\,\frac {\delta Q^2}{{q^+}^2} \left(\frac{\Theta_{e \bar e}} {Q_{e}^2} - \frac{\Theta_{\bar e e}} {Q_{\bar e}^2}\right) \right] \nonumber\\ &-& \frac{\alpha}{4\pi^2} \langle \gamma^\mu\gamma^\nu\rangle_{an} \left[ g_{\mu\nu} \left(\frac{\Theta_{ab}} {M_a^2} + \frac{\Theta_{ba}} {M_b^2} \right) - \eta_\mu\eta_\nu\,\frac{\delta M^2} {{p^+}^2} \left(\frac{\Theta_{ab}} {M_a^2} - \frac{\Theta_{ba}} {M_b^2} \right) \right] \,.\label{eq:56}\end{eqnarray} with the generator of unitary transformation \begin{eqnarray} \eta(l) &=& -\frac{1}{D}\left(\frac{d{\rm ln}\,f(D;l)}{dl}\right)g(l) \,.\label{eq:gen}\end{eqnarray} where $g(l)$ is the coupling constant as a function of flow parameter $l$, and $f(D;l)$ is the cutoff function specified below. In Eq.~(\ref{eq:56}) subscript $ex$ refers to the exchange part, and $an$ to the annihilation part. The null vector $\eta^{\mu}$ has components $(\eta^+,\vec\eta_\perp, \eta^-)=(0,\vec{0},2)$ and is specific to the light-front calculations. The light-front metric tensor is denoted by $g_{\mu\nu}$. The current-current tensors in the two channels are \begin{eqnarray} \langle \gamma^\mu\gamma^\nu\rangle_{ex} &=& \frac{(\overline u(p_1,\lambda_1) \gamma^\mu u(p'_1,\lambda'_1))\, (\overline v(p'_2,\lambda'_2) \gamma^\nu v(p_2,\lambda_2))} {\sqrt{x x' (1-x) (1-x')} } \,,\nonumber\\ \langle \gamma^\mu\gamma^\nu\rangle_{an} &=& \frac{(\overline u(p_1,\lambda_1)\gamma^\mu v(p_2,\lambda_2))\, (\overline v(p'_2,\lambda'_2)\gamma^\nu u(p'_1,\lambda'_1))} {\sqrt{x x' (1-x) (1-x')} } \,,\label{eq:r94}\end{eqnarray} where the fermion momenta were defined after Eq.~(\ref{eq:i3}). The remaining definitions are as follows. The energy differences along the electron and the positron line, \begin{eqnarray} D_{e} &=& {p'_1}^- - p_1^{-} - (p'_1-p_1)^- \,,\nonumber\\ D_{\bar e} &=& p_2^{-} - {p'_2}^- - (p_2-p'_2)^- \,.\label{eq:40a} \end{eqnarray} respectively, have a simple relation to the (Feynman-) 4-momentum transfers along the two lines \begin{eqnarray} Q_{e}^2 &=& -(p'_1-p_1)^2 = -q^+D_{e} \,,\nonumber\\ Q_{\bar e}^2 &=& -(p_2-p'_2)^2 = -q^+D_{\bar e} \,.\end{eqnarray} Since the Feynman-momentum transfer $Q$ is more physical quantity than the energy difference, we will use the former as far as possible. In fact, in our formulae we make use of the {\em mean-square momentum transfer} and the {\em mean-square difference}, \begin{eqnarray} Q^2 &=& {1\over 2}(Q_{e}^2+Q_{\bar e}^2) = -{q^+\over 2}(D_{e}+D_{\bar e}) \,,\nonumber\\ \delta Q^2 &=& {1\over 2}(Q_{e}^2-Q_{\bar e}^2) = -{q^+\over 2}(D_{e}-D_{\bar e}) \,,\label{eq:r15}\end{eqnarray} respectively. The dependence of the effective interaction Eq.~(\ref{eq:56}) on the cutoff function $f(D;l)$ is carried by the factor \begin{equation} \Theta(D_{e},D_{\bar e}) = - \int_{0}^\infty\!dl'\, \frac{d f (D_{e};l')}{dl'} f (D_{\bar e};l') \equiv \Theta_{e\bar e} \,,\label{eq:r16}\end{equation} which is asymmetric in the arguments but which satisfies \begin{equation} \Theta(D_{e},D_{\bar e})+\Theta(D_{\bar e},D_{e}) = \Theta_{e\bar e} + \Theta_{\bar e e} = 1 \,.\label{eq:ii18a}\end{equation} The latter combination obeys \begin{eqnarray} \frac{\Theta_{e \bar{e}}} {Q_{e}^2}+ \frac{\Theta_{\bar{e} e}} {Q_{\bar{e}}^2} &=& \frac {Q^2} {Q_{e}^2 Q_{\bar{e}}^2} \left( 1 - \frac{\delta Q^2}{Q^2} \left(\Theta_{e\bar{e}} - \Theta_{\bar{e}e} \right) \right) \nonumber\\ \frac{\Theta_{e\bar{e}}} {Q_{e}^2}- \frac{\Theta_{\bar{e}e}}{Q_{\bar{e}}^2} &=& - \frac {\delta Q^2} {Q_{e}^2 Q_{\bar{e}}^2} \left( 1 - \frac{ Q^2}{\delta Q^2} \left(\Theta_{e\bar{e}} - \Theta_{\bar{e}e} \right) \right) \,.\label{eq:ii18b}\end{eqnarray} that we use further. For the annihilation term we define the energy differences as \begin{eqnarray} D_a &=& {p'_1}^- + {p'_2}^- - (p'_1+p'_2)^- \nonumber\\ D_b &=& p_1^{-} + p_2^{-} - (p_1+p_2)^- \,.\label{eq:40b} \end{eqnarray} They are related to the 4-momentum $p^\mu$ of the photon and to the free invariant mass-squares of the initial and final states \begin{eqnarray} M_a^2 &=& (p'_1+p'_2)^2 = p^+ D_a \,,\nonumber\\ M_b^2 &=& (p _1+p _2)^2 = p^+ D_b \,,\end{eqnarray} as well as to their mean and difference \begin{eqnarray} M^2 &=& {1\over 2}(M_a^2 + M_b^2)= {p^+\over 2}(D_a+D_b) \,,\nonumber\\ \delta M^2 &=& {1\over 2}(M_a^2 - M_b^2)= {p^+\over 2}(D_a-D_b) \,,\label{eq:r16a}\end{eqnarray} respectively. Effective interaction~(\ref{eq:56}) includes two different Lorentz structures: $g_{\mu\nu}$ part insures the Bohr spectrum and is responsible for the spin splittings; $\eta_{\mu}\eta_{\nu}$ term is diagonal in spin space and vanishes for real processes, i.e. on mass shell with $\delta Q^2=0$, making the effective interaction to coincide with the Tamm-Dancoff approximation ~\cite{previous}. The explicit $x$-dependence in the denominator of Eq.~(\ref{eq:r94}) looks like the only remnant of the light-front formulation; all other quantities are Lorentz scalars. One can absorb this dependence by redefining the wave function in the integral equation Eq.~(\ref{eq:r96}). We introduce instead of Jacobi momentum $(x,\vec{\kappa}_{\perp})$ the three momentum in the center of mass frame $\vec{p}=(p_z,\vec{\kappa}_{\perp})$ as follows \begin{eqnarray} x = \frac{1}{2}\left(1+\frac{p_z}{\sqrt{{\vec p}^{\, 2}+m^2}} \right) \,,\label{eq:eq3}\end{eqnarray} where the Jacobian of this transformation $dx/dp_z$ is \begin{eqnarray} J = \frac{1}{2}\frac{\vec{\kappa}_{\perp}^2+m^2} {({\vec p}^{\, 2}+m^2)^{3/2}} = 2\frac{x(1-x)}{E} \,,\label{eq:eq4}\end{eqnarray} and it holds in this frame \begin{eqnarray} x(1-x) &=& \frac{1}{4}\frac{\vec{\kappa}_{\perp}^2+m^2}{{\vec p}^{\, 2}+m^2} \,,\nonumber\\ E &=& \sqrt{{\vec p}^{\, 2}+m^2} \,.\end{eqnarray} The connection between `old' and `new' wave functions and the interaction matrix elements are \begin{eqnarray} \langle x,\vec{\kappa}_{\perp} \vert \psi \rangle &=& \frac{\langle \vec{p} \vert \psi^\prime \rangle}{\sqrt{x(1-x)}} \,,\nonumber\\ \langle x,\vec \kappa_{\!\perp} \vert V_{\rm eff} \vert x^\prime,\vec \kappa_{\!\perp}^\prime \rangle &=& \frac{ \langle \vec {p} \vert V_{\rm eff}^\prime \vert {\vec p}^{\: \prime} \rangle} {\sqrt{x(1-x)x'(1-x')}} \,.\end{eqnarray} Integral equation~(\ref{eq:r96}), \begin{eqnarray} M^2\ \langle \vec{p}; \lambda_1, \lambda_2 \vert \psi^\prime \rangle &=& 4 ({\vec p}^{\, 2}+m^2) \ \langle \vec p; \lambda_1, \lambda_2 \vert \psi^\prime \rangle \nonumber\\ &+& \sum _{ \lambda_1^\prime,\lambda_2^\prime} \!\int_D\! \frac{d^3 {\vec p}^{\: \prime}}{2E}\, \,\langle \vec p; \lambda_1, \lambda_2 \vert V_{\rm eff}^\prime \vert {\vec p}^{\: \prime}; \lambda_1^\prime, \lambda_2^\prime\rangle \ \langle {\vec p}^{\: \prime}; \lambda_1^\prime,\lambda_2^\prime \vert \psi^\prime \rangle \label{eq:r96a}\end {eqnarray} is written in a rotationally covariant form. \section{Rotational invariance} \label{sec:3} Integral equation~(\ref{eq:r96}) has rotationally covariant but still not rotationally invariant form because of the interaction kernel $\tilde{V}_{eff}$, written in the light-front frame. Let us extract the part of interaction, which has manifestly rotational symmetry. Quite generally $\Theta$ factor Eq.~(\ref{eq:r16}) is a function of the ratio of its two arguments. Therefore, making use of Eq.~(\ref{eq:ii18b}), the effective interaction Eq.~(\ref{eq:56}) is given as \begin{eqnarray} V_{\rm eff} = - \frac{\alpha}{4\pi^2} \langle\gamma^\mu\gamma^\nu\rangle_{ex} B_{\mu\nu} - \frac{\alpha}{4\pi^2} \langle\gamma^\mu\gamma^\nu\rangle_{an} C_{\mu\nu} \,,\label{eq:inter}\end{eqnarray} where the exchange part is defined \begin{eqnarray} B_{\mu\nu} &=& \frac{g_{\mu\nu}}{Q^2} + \left( \frac{g_{\mu\nu}}{Q^2} - \frac{\eta_{\mu}\eta_{\nu}}{{q^+}^2} \right) \frac{\xi^2-\xi\vartheta(\xi)}{1-\xi^2} \nonumber\\ \vartheta(\xi) &=& \Theta_{e \bar{e}} - \Theta_{\bar{e} e} \quad,\quad \xi = \frac{\delta Q^2}{Q^2} \,,\label{eq:eq1}\end{eqnarray} and the annihilation part \begin{eqnarray} C_{\mu\nu} &=& \frac{g_{\mu\nu}}{M^2} + \left( \frac{g_{\mu\nu}}{M^2} - \frac{\eta_{\mu}\eta_{\nu}}{{p^+}^2} \right) \frac{\beta^2-\beta\chi(\beta)}{1-\beta^2} \nonumber\\ \chi(\beta) &=& \Theta_{a b} - \Theta_{b a} \quad,\quad \beta = \frac{\delta M^2}{M^2} \,.\label{eq:eq2}\end{eqnarray} The terms in the effective interaction proportional to $\vartheta(\xi)$, $\chi(\beta)$ depend explicitly on the choice of cut-off function and arise from $l$-ordering of the generator in the operator of unitary transform~\cite{previous}. Explicit form of the effective interaction with different cut-off functions is given in Appendix A. Define energy denominators in equation~(\ref{eq:eq2}). Due to the three-momentum conservation on the light-front, $p_1+p_2=p'_1+p'_2$ for longitudinal and transversal components, one has $D_e-D_{\bar e}=D_a-D_b$, where the energy denominators $D_k$ in both channels are given in Eq.~(\ref{eq:40a}) and Eq.~(\ref{eq:40b}). Therefore \begin{eqnarray} \delta Q^2 = \left( - \frac{q^+}{p^+} \right)\delta M^2 \,.\label{eq:delta}\end{eqnarray} where $\delta M^2$ -- the total energy difference between initial and final states shows the 'off-shellness` of process. Using the parametrization Eq.~(\ref{eq:eq3}) one has for the energy denominators \begin{eqnarray} Q^2 &=& {\vec q}^{\, 2} -p_z p'_z\frac{(M_a-M_b)^2}{M_aM_b} \nonumber\\ \delta Q^2 &=& - \left( \frac{p'_z}{M_a} - \frac{p_z} {M_b} \right)\delta M^2 \nonumber\\ M_a^2 &=& 4({\vec p}^{\: \prime 2}+m^2) \nonumber\\ M_b^2 &=& 4({\vec p}^{\: 2}+m^2) \,,\label{eq:note}\end{eqnarray} where $q=p'-p=(q_z,q_{\perp})$ is the three-momentum transfer of the photon, and the relations between mean-squared and difference momenta and corresponding energy differences are given in Eq.~(\ref{eq:r15}) and Eq.~(\ref{eq:r16}) for exchange and annihilation channels, respectively. The second term in Eq.~(\ref{eq:eq2}) is obviously not rotational invariant. For the real processes, $\delta M^2=0$, the second term vanishes for both channels, and the effective interaction is independent on the cutoff function and coincide with the result of Tamm-Dancoff approximation \begin{eqnarray} V_{\rm eff} &=& -\frac{\alpha}{4\pi^2} \frac{\langle\gamma^\mu \gamma_\mu\rangle}{{\vec q}^{\,2}} \,,\label{eq:td}\end{eqnarray} and similarly in annihilation channel. The same holds to the leading order of nonrelativistic expansion ${\vec p}^{\, 2}/m^2 \ll 1$ ~\cite{previous}. Estimate current-current term in exchange channel, Eq.~(\ref{eq:r94}), which defines nominator of the interaction Eq.~(\ref{eq:td}). We work in the Lepage-Brodsky convention for the spinors~\cite{LeBr} \begin{eqnarray} u(p,\lambda) &=& \frac{2}{p^+} ( p^+ +\beta m + \vec{\alpha}_{\perp}\vec{p}_{\perp} ) \Lambda_{+} \chi_{\lambda} \nonumber\\ &=& \frac{2}{\sqrt{N}} ( E +\beta m + \vec{\alpha}\vec{p} ) \Lambda_{+} \chi_{\lambda} \,,\label{eq:spinor}\end{eqnarray} and similarly for $v(p,\lambda)$ with the change $m\rightarrow -m$ and $\chi_{\lambda}\rightarrow \chi_{-\lambda}$ in the above formula. The second expression for spinor holds quite in general for the solution of Dirac equation, where $p^0=E=\sqrt{{\vec p}^{\, 2}+m^2}$ and the normalization factor $N=E+p_z$. Here $\beta=\gamma^0$, $\vec{\alpha}=\gamma^0\vec{\gamma}$, $\Lambda_{+}=1/2(1+\alpha^3)$ is the projection operator ~\cite{bpp97}, and the spinor \begin{displaymath} \chi_{\lambda} = {\xi_{\lambda} \choose 0} \,,\label{eq:spinor1}\end{displaymath} is defined through the usual two-component spinors \begin{displaymath} \xi_{\uparrow} ={1 \choose 0} \quad; \quad \xi_{\downarrow}{0 \choose 1} \,,\label{eq:spinor2}\end{displaymath} Using the explicit representation for $\gamma$ matrices and projection operators $\Lambda_{+}, \Lambda_{-}$ and relations between them ~\cite{bpp97}, one has \begin{eqnarray} \bar{u}(p,\lambda)\gamma^0 u(p^\prime,\lambda^\prime) &=& \frac{1}{\sqrt{NN^\prime}}\xi_{\lambda}^+ ( (E+p_z)(E^\prime+p_z^\prime) +(\vec p {\vec p}^{\: \prime}) +i[\vec p \times {\vec p}^{\: \prime}]\vec \sigma \nonumber\\ &+& i[\vec p \times \vec \sigma]_z(p_z^\prime+m) -i[{\vec p}^{\: \prime}\times \vec \sigma]_z(p_z+m) +m^2-p_zp_z^\prime ) \xi_{\lambda^\prime} \nonumber\\ \bar{u}(p,\lambda)\gamma^i u(p^\prime,\lambda^\prime) &=& \frac{1}{\sqrt{NN^\prime}}\xi_{\lambda}^+ ( (E+p_z) (p^{\prime\, i} +i[{\vec p}^{\: \prime} \times \vec \sigma]^{i}) + (E^\prime+p^\prime_z)(p^{i} -i[\vec p \times \vec \sigma]^{i}) \nonumber\\ &+& \delta^{iz} ( EE^\prime-m^2-(\vec p {\vec p}^{\: \prime}) -i[\vec p \times {\vec p}^{\: \prime}]\vec{\sigma} \nonumber\\ &+& i[\vec p \times \vec \sigma]^{i}(E^\prime -m) -i[{\vec p}^{\: \prime}\times \vec \sigma]^{i}(E -m) ) \nonumber\\ &+& i\varepsilon^{ij}\sigma^{j} ( (E+p_z)(m+p_z^\prime)-(E^\prime+p_z^\prime)(m+p_z) ) ) \xi_{\lambda^\prime} \,,\label{eq:a}\end{eqnarray} where $i=1,2,3$; $p=(p_z,\vec{\kappa}_{\perp})$ is the three momentum and $\varepsilon^{ij}=\varepsilon^{ijz}$, $\varepsilon^{12}=-\varepsilon^{21}=1$. Introducing three-momentum transfer and its mean \begin{eqnarray} \vec q &=& {\vec p}^{\: \prime}-\vec p \nonumber\\ \vec k &=& \frac{1}{2}({\vec p}^{\: \prime}+\vec p) \,,\label{}\end{eqnarray} where $4(\vec q\vec k)=\delta M^2=2(E^2-E^{\prime\, 2})$ with $\delta M^2$ defined above, Eq.~(\ref{eq:a}) is written \begin{eqnarray} \bar{u}(p,\lambda)\gamma^0 u(p^\prime,\lambda^\prime) &=& \frac{1}{\sqrt{NN^\prime}}\xi_{\lambda}^+ ( (E+k_z)(E^\prime +k_z) +\vec k^2-{\vec q}^{\, 2}/4 -i[\vec q \times \vec k]\vec \sigma \nonumber\\ &-& i[\vec q \times \vec \sigma]_z (k_z+m) +q_z((E-E^\prime)/2 +i[\vec k \times \vec \sigma]_z) +m^2-k_z^2 ) \xi_{\lambda^\prime} \nonumber\\ \bar{u}(p,\lambda)\gamma^i u(p^\prime,\lambda^\prime) &=& \frac{1}{\sqrt{NN^\prime}}\xi_{\lambda}^+ ( ((E+E^\prime)/2+k_z) (2k^{i} +i[\vec q \times\vec \sigma]^{i}) \nonumber\\ &+& ((E-E^\prime) -q_z) (q^{i}/2 +i[\vec k \times\vec \sigma]^{i}) \nonumber\\ &+& \delta^{iz} ( EE^\prime-m^2-(\vec{k}^2-{\vec q}^{\, 2}/4) +i[\vec q \times\vec k]\vec{\sigma} \nonumber\\ &-& i[\vec q\times \vec \sigma]^{i} ((E+E^\prime)/2-m) - i[\vec k \times\vec \sigma]^{i} (E-E^\prime) ) \nonumber\\ &+& i\varepsilon^{ij}\sigma^{j} ((k_z+m)(E-E^\prime)+q_z((E+E^\prime)/2-m)) ) \xi_{\lambda^\prime} \,,\label{eq:b}\end{eqnarray} where no approximations are done so far. Excluding the overall normalization factor the first lines in Eq.~(\ref{eq:b}) for scalar and vector current terms contain rotationally invariant parts (except terms proportional to $k_z$), which coincide with the corresponding expressions when making use of Bjorken-Drell convention for spinors ~\cite{MePa}. \footnote{ Spinors as used by Bjorken-Drell are \begin{eqnarray} u(p,\lambda) &=& \frac{1}{\sqrt{N}} ( E +\beta m + \vec{\alpha}\vec{p} ) \chi_{\lambda} \,,\label{eq:fn1}\end{eqnarray} where $N=E+m$ and $\chi_{\lambda}$ is defined in the main text. The correspoding expressions for current terms are \begin{eqnarray} \bar{u}(p,\lambda)\gamma^0 u(p^\prime,\lambda^\prime) &=& \frac{1}{\sqrt{NN^\prime}}\xi_{\lambda}^+ ( (E+m)(E^\prime+m) +\vec{k}^2-\vec{q}^{\, 2}/4 -i[\vec q \times \vec k]\vec \sigma ) \xi_{\lambda^\prime} \nonumber\\ \bar{u}(p,\lambda)\gamma^i u(p^\prime,\lambda^\prime) &=& \frac{1}{\sqrt{NN^\prime}}\xi_{\lambda}^+ ( ((E+E^\prime)/2+m) (2k^{i} +i[\vec q \times \vec \sigma]^{i}) \nonumber\\ &+& (E-E^\prime) (q^{i}/2 +i[\vec k \times \vec \sigma]^{i}) ) \xi_{\lambda^\prime} \,,\label{eq:fn2}\end{eqnarray} For the energy conserving process this expression was obtained in ~\cite{MePa}. } Merkel et.al.~\cite{MePa} showed, that as far as the energy is conserved, this part gives rise to familiar spin dependent forces. The rest terms in Eq.~(\ref{eq:b}) are obviously not rotationally invariant, particularly when the spacial rotations are performed perpendicular to the $z$-axis. Expanding expression Eq.~(\ref{eq:b}) to the second order in $|\vec p|/m \ll 1$ and performing the unitary transformation in spin space, Brisudova et.al.~\cite{BrPe2} obtained Breit-Fermi spin-spin and tensor interactions. It seems to be impossible to reproduce full set of Breit-Fermi terms from the second order effective interaction in the light-front gauge. Also it is complicated to cover rotational symmetry on the level of light-front effective Hamiltonians without additional approximations are done. We use directly the effective electron-positron interaction Eq.~(\ref{eq:inter}) for numerical calculations of positronium spectrum. We aim to get fine structure and to investigate rotational symmetry on the level of spectrum. The impact of different cutoff functions is also considered. The results of these calculations are presented in the next section. \begin{table} \begin{tabular}{c||c||c||c||c||c||c} $n$ & Term & $B_{ETPT}$ & $B_E$ & $B_G^{\eta}$& $B_G$ & $B_S$ \cr \hline \hline 1 & $1^1S_0$ & 1.118125 & 1.049550 & 1.101027 & 1.026170 & 0.920921 \\ 2 & $1^3S_1$ & 0.998125 & 1.001010 & 1.049700 & 0.981969 & 0.885347 \\ 3 & $2^1S_0$ & 0.268633 & 0.260237 & 0.266490 & 0.260642 & 0.242607 \\ 4 & $2^3S_1$ & 0.253633 & 0.253804 & 0.259506 & 0.254765 & 0.234312 \\ 5 & $2^1P_1$ & 0.253633 & 0.257969 & 0.263056 & 0.257664 & 0.237611 \\ 6 & $2^3P_0$ & 0.261133 & 0.267070 & 0.273826 & 0.266563 & 0.243075 \\ 7 & $2^3P_1$ & 0.255508 & 0.259667 & 0.265412 & 0.260127 & 0.238135 \\ 8 & $2^3P_2$ & 0.251008 & 0.255258 & 0.260345 & 0.255498 & 0.236383 \end{tabular} \caption{Binding coefficients, $B_n=4 (2-M_n)/\alpha^2$ ($\alpha=0.3$), for the lowest modes of the positronium spectrum at $J_z=0$ for the equal time perturbation theory up to order $\alpha^4$ ($B_{ETPT}$) compared to our calculations with exponential ($B_E$), Gaussian ($B_G$) and sharp ($B_S$) cutoffs. $B_G$ is obtained using only $g_{\mu\nu}$ part of interaction; for $B_G^{\eta}$ $'\eta_\mu\eta_\nu`$ term is included. Exchange channel is considered.} \end{table} \begin{table} \begin{tabular}{c||c||c||c||c||c} $n$ & Term & $B_E$ & $B_G^{\eta}$& $B_G$ & $B_S$ \cr \hline \hline 1 & $1^1S_0$ & 1.049550 & 1.101270 & 1.026170 & 0.920921 \\ 2 & $1^3S_1$ & 0.936800 & 0.978018 & 0.921847 & 0.834004 \\ 3 & $2^1S_0$ & 0.260237 & 0.266490 & 0.260642 & 0.242624 \\ 4 & $2^3S_1$ & 0.255292 & 0.260383 & 0.255615 & 0.234338 \\ 5 & $2^1P_1$ & 0.257969 & 0.263056 & 0.257664 & 0.236383 \\ 6 & $2^3P_0$ & 0.267090 & 0.273847 & 0.266626 & 0.243075 \\ 7 & $2^3P_1$ & 0.259667 & 0.265412 & 0.260127 & 0.237611 \\ 8 & $2^3P_2$ & 0.245615 & 0.250821 & 0.247091 & 0.230901 \end{tabular} \caption{Binding coefficients, $B_n=4 (2-M_n)/\alpha^2$ ($\alpha=0.3$), for the lowest modes of the positronium spectrum at $J_z=0$ for our calculations with exponential ($B_E$), Gaussian ($B_G$) and sharp ($B_S$) cutoffs. $B_G^{\eta}$ includes $'\eta_\mu\eta_\nu`$ term in exchange channel; $B_G$ does not. Exchange and annihilation channels are considered.} \end{table} \begin{table} \begin{tabular}{c|| c|| c|| c|| c} $n$ &Term & $\delta B_E$ & $\delta B_G$ & $\delta B_S$ \\ \hline \hline\\[-8pt] 2 & $1^3S_1$ & 6.30 $10^{-4}$ & 1.76 $10^{-3}$ & 1.18 $10^{-3}$ \\ 4 & $2^3S_1$ & 8.40 $10^{-5}$ & 1.77 $10^{-4}$ & 9.0 $10^{-5}$ \\ 5 & $2^1P_1$ & -1.30 $10^{-5}$ & -7.47 $10^{-4}$ &-9.1 $10^{-5}$ \\ 7 & $2^3P_1$ & -4.08 $10^{-4}$ & -4.08 $10^{-4}$ & 1.4 $10^{-4}$ \\ 8 & $2^3P_2$ & 5 $10^{-6}$ & -7.7 $10^{-5}$ & 4.15 $10^{-4}$ \end{tabular} \caption{Difference in the corresponding energy levels between $J_z\!=\!0$ and $J_z\!=\!1$ states for exponential ($\delta B_E$), Gaussian ($\delta B_G$) and sharp ($\delta B_S$) cutoffs. Exchange is channel is considered.} \end{table} \begin{table} \begin{tabular}{c|| c|| c|| c|| c} $n$ & Term & $\delta B_E$ & $\delta B_G$ & $\delta B_S$ \\ \hline \hline\\[-8pt] 2 & $1^3S_1$ & -1.411 $10^{-3}$ & -7.86 $10^{-4}$ &-1.65 $10^{-3}$ \\ 4 & $2^3S_1$ & -4.1 $10^{-5}$ & -4.0 $10^{-5}$ &-1.15 $10^{-4}$ \\ 5 & $2^1P_1$ & -6.4 $10^{-5}$ & -6.52 $10^{-4}$ &-4.60 $10^{-4}$ \\ 7 & $2^3P_1$ & -4.69 $10^{-4}$ & -4.74 $10^{-4}$ &-1.40 $10^{-4}$ \\ 8 & $2^3P_2$ & -1.96 $10^{-4}$ & -1.36 $10^{-4}$ &-2.44 $10^{-4}$ \end{tabular} \caption{Difference in the corresponding energy levels between $J_z\!=\!0$ and $J_z\!=\!1$ states for exponential ($\delta B_E$), Gaussian ($\delta B_G$) and sharp ($\delta B_S$) cutoffs. Exchange and annihilation channels are considered.} \end{table} \section{Mass spectrum of positronium} \label{sec:6.3} We solve the integral equation~(\ref{eq:r96}), with interaction kernel given in Eq.(\ref{eq:56}), for positronium spectrum numerically. Effective interaction with different choice of cutoffs is sumarized in Appendix A. In polar coordinates the light-front variables are $(\vec \kappa_{\perp};x)=(\kappa_{\perp},\varphi;x)$; therefore the matrix elements of the effective interaction Eq.(\ref{eq:56}) depend on the angles $\varphi$ and $\varphi^{\prime}$, i.e. $\langle x,\kappa_{\perp},\varphi;\lambda_1,\lambda_2|V_{\rm eff}| x',\kappa'_{\perp},\varphi';\lambda'_1,\lambda'_2\rangle$. In order to introduce the spectroscopic notation for positronium mass spectrum we integrate out the angular degree of freedom, $\varphi$, introducing a discrete quantum number $J_z=n$, $n\in {\bf Z}$ (actually for the annihilation channel only $|J_z|\leq 1$ is possible), \begin{eqnarray} && \hspace{-1.5cm} \langle x, \kappa_{\perp}; J_z, \lambda_1, \lambda_2 |\tilde{V}_{\rm eff}| x',\kappa'_{\perp};J'_z,\lambda'_1,\lambda'_2\rangle \nonumber\\ &=& \frac{1}{2\pi}\int_0^{2\pi}d\varphi {\rm e}^{-iL_z\varphi} \int_0^{2\pi}d\varphi'{\rm e}^{iL'_z\varphi'} \langle x, \kappa_{\perp}, \varphi; \lambda_1, \lambda_2 |V_{\rm eff}| x',\kappa'_{\perp},\varphi';\lambda'_1,\lambda'_2\rangle \nonumber\\ && \,\label{eq:r43}\end{eqnarray} where $L_z=J_z-S_z$; $S_z=\frac{\lambda_1}{2}+\frac{\lambda_2}{2}$ and the states can be classified (strictly speaking only for rotationally invariant systems) according to their quantum numbers of total angular momentum $J$, orbit angular momentum $L$, and total spin $S$. Definition of angular momentum operators in light-front dynamics is problematic because they include interactions. The matrix elements of the effective interaction before integrating over the angles, $ \langle x, \kappa_{\perp}, \varphi; \lambda_1, \lambda_2 |V_{\rm eff}| x',\kappa'_{\perp},\varphi';\lambda'_1,\lambda'_2\rangle$, and after the integration inroducing the total momentum, $J_z$, $ \langle x, \kappa_{\perp}; J_z, \lambda_1, \lambda_2|\tilde{V}_{\rm eff}| x',\kappa'_{\perp};J'_z,\lambda'_1,\lambda'_2\rangle$ for different cutoff functions are given in the exchange and annihilation channels in Appendices B and C, respectively. Now we proceed to solve for the positronium spectrum in all sectors of $J_z$. For this purpose we formulate the light-front integral equation Eq.~(\ref{eq:r96}) in the form where the integral kernel is given by the effective interaction for the total momentum $J_z$, Eq.~(\ref{eq:r43}). After the change of variables Eq.~(\ref{eq:eq3}) we parametrize $ \vec{p}=(\vec{\kappa}_{\perp},p_z)= (\mu\sin\theta\cos\varphi,\mu\sin\theta\sin\varphi,\mu\cos\theta)$. The Jacobian of this transformation Eq.~(\ref{eq:eq4}) is given \begin{eqnarray} && J=\frac{1}{2} \frac{m^2+\mu^2 \sin^2\theta}{(m^2+\mu^2)^{3/2}} \,.\label{r47}\end{eqnarray} \newpage One obtaines then the integral equation \begin{eqnarray} && (M_n^2-4(m^2+\mu^2)) \tilde{\psi}_n(\mu,\cos\theta;J_z, \lambda_1, \lambda_2) \nonumber\\ &+& \sum_{J'_z,\lambda'_1,\lambda'_2} \int_{D}d\mu'\int_{-1}^{+1}d\cos\theta' \frac{\mu^{'2}}{2} \frac{m^2+\mu^{'2}(1-\cos^2\theta')}{(m^2+\mu^{'2})^{3/2}} \nonumber\\ &\times& \langle\mu, \cos\theta; J_z, \lambda_1, \lambda_2 |\tilde{V}_{\rm eff}| \mu',\cos\theta';J'_z,\lambda'_1,\lambda'_2\rangle \tilde{\psi}_n (\mu',\cos\theta';J'_z,\lambda'_1,\lambda'_2)=0 \,.\label{eq:r48}\end{eqnarray} The integration domain $D$, defined in Eq.~(\ref{eq:domain}), is given now by $\mu\in [0;\frac{\Lambda}{2}]$. Neither $L_z$ nor $S_z$ are good quantum numbers; therefore we set $L_z=J_z-S_z$. The integral equation Eq.~(\ref{eq:r48}) is used to calculate positronium mass spectrum numerically. Note, that if one succeeds to integrate out the angular degrees of freedom for the effective interaction Eq.~(\ref{eq:r43}) analytically, one has $2$-dimensional integration in Eq.~(\ref{eq:r48}) instead of $3$-dimensional one in the original integral equation~(\ref{eq:r96}) to perform numerically. We use the numerical code ~\cite{TrPa2}, worked out by Uwe Trittmann for the similar problem ~\cite{TrPa}. This code includes for the numerical integration the Gauss-Legendre algorithm (Gaussian quadratures). To improve the numerical convergence the technique of Coulomb counterterms is included. The problem has been solved for all components of the total angular momentum, $J_z$. Positronium spectrum is mainly defined by the Coulomb singularity \begin{eqnarray} \vec q \longrightarrow 0 \,,\label{eq:limit1}\end{eqnarray} which is an integrable one analytically and also, by use of technique of Coulomb counterterms, numerically. In this region $\delta Q^2\rightarrow 0$ and the energy denominator $Q^2\rightarrow {\vec q}^{\, 2}$ Eq.~(\ref{eq:note}), giving rise to the leading order Coulomb behavior for the effective interaction Eq.~(\ref{eq:td}), independent on the cutoff function. We use therefore standard Coulomb counterterms, introduced for the Coulomb problem Eq.~(\ref{eq:td}) ~\cite{TrPa,TrPa2}, in the case of all cutoffs. Basing also on the argument Eq.~(\ref{eq:limit1}), we expect the same pattern of levels for different cutoffs, that we prove numerically to be true. Another important limiting case to study effective interaction Eq.~(\ref{eq:inter}), namely its exchange part, is the collinear limit \begin{eqnarray} q^+ \longrightarrow 0 \,,\label{eq:limit2}\end{eqnarray} that is special for light-front calculations. Because of Eq.~(\ref{eq:delta}) the variable $\xi^2 \sim q^{+\, 2}$, resulting for the $'\eta_\mu\eta_\nu'$ part of effective interaction to be \begin{eqnarray} \frac{\eta_\mu\eta_\nu}{q^{+\, 2}}\xi^2(1-\vartheta^{\prime}(0)) \,,\label{eq:}\end{eqnarray} which is finite in this limit. This is true for the regular cutoff functions, as in the case of exponential and gaussian cutoffs, where the derivative $d\vartheta(0)/d\xi$ is well defined. For sharp cutoff this condition is not fulfilled, and the effective interaction contains the $1/q^+$ type of singularity in this case (see Appendix A). We do not associate any physics with this singularity, considering it as a consequence of artificial choice of cutoff, which corresponds to singular generator of unitary transformation Eq.~(\ref{eq:gen}). We omit the $'\eta_\mu\eta_\nu'$ term in exchange channel for sharp cutoff in numerical calculations. We argued that the region of Coulomb singularity, and hence $'g_{\mu\nu}'$ part of effective interaction, determines mainly the positronium spectrum. However, including $'\eta_\mu\eta_\nu'$ part for gaussian cutoff shifts all levels as a whole of about $5-7\%$, since this part is diagonal in spin space (Appendix B), and improves the data to be near the result obtained in covariant equal time calculations (Table $1$). Presumably, it is necessary to take into account $'\eta_\mu\eta_\nu'$ term in exchange channel also for sharp cutoff after the proper regularization of infrared longitudinal divergences is done. We place the results of calculations for three different cutoffs, performed in exchange and including both exchange and annihilation channels, in Tables $1$ and $2$, respectively. The corresponding set of figures is presented in Fig.$1$ and Fig.$2$. We get the ionization threshold at $M^2\sim 4m^2$, the Bohr spectrum, and the fine structure. Including annihilation part increases the splittings twice as large for the lowest multiplets. As one can see from presents figures, certain mass eigenvalues at $J_z=0$ are degenerate with certain eigenvalues at other $J_z$ to a very high degree of numerical precision. As an example, consider the second lowest eigenvalue for $J_z=0$. It is degenerate with the lowest eigenvalue for $J_z=\pm 1$, and can thus be classified as a member of the triplet with $J=1$. Correspondingly, the lowest eigenvalue for $J_z=0$ having no companion can be classified as the singlet state with $J=0$. Quite in general one can interpret degenerate multiplets as members of a state with total angular momentum $J=2J_{z,max}+1$. One can get the quantum number of total angular momentum $J$ from the number of degenerate states for a fixed eigenvalue $M_n^2$. One can make contact with the conventional classification scheme $^{2S+1}L_{J}^{J_z}$, as indicated in Tables $1-2$. Such pattern of spectrum is driven by rotational invariance. To trace rotational symmetry we calculate the difference of energy levels between $J_z=0$ and $J_z=1$ states for the lowest multiplets. The data are given for exchange and including annihilation channnel in Tables $3$ and $4$, respectively. Annihilation part makes corresponding states practically degenerate (see Tables $4$ and Figure $2$). \section{Conclusion} \label{sec:6.4} The numerical solution of positronium bound state problem, with the effective electron-positron interaction obtained by the flow equations, is presented. No approximations along numerical procedure are done. Concerning the spin-splittings the best agreement with covariant calculations is obtained for gaussian cutoff, the worst results are for sharp cutoff. Rotational invariance is traced on the level of spectrum by studing the degree of degeneracy of corresponding states with the same total momentum but different projection $J_z$ in the multiplet. Again, better results are obtained for exponential and gaussian cutoff functions than for sharp cutoff. This suggests, that smooth cutoff functions are preferable to perform calculations. Including annihilation channel improves the extend of degeneracy. For the sharp cutoff the lowest multiplet is placed higher than the one in case of exponential and gaussian cutoffs. The reason is in disregarding the infrared divergent part, which is diagonal in spin space and shifts the spectrum as a whole down. The question how to regularize this part and include it in mass spectrum calculations should be considered. Generally, the impact of the different choice of cutoff functions on the spectrum is small. In this work we solve the bound state integral equation for the one fixed integration interval. Integration domain introduces the ultraviolet-cutoff dependence of invariant mass squared $M^2(\Lambda)$, that reflects renormalization group properties of the effective coupling constant. We leave this question for the future study. \newpage \begin{figure}[thbp] \centerline{\epsfxsize=\textwidth \epsfbox{noanf.eps}} \caption{The invariant mass-squared spectrum $M_i^2$ for positronium versus the projection of the total spin, $J_z$, excluding annihilation with exponential, Gaussian and sharp cutoffs. The number of integration points is $N_1=N_2=21$.} \label{fig:1} \end{figure} \begin{figure}[thbp] \centerline{\epsfxsize=\textwidth \epsfbox{annif.eps}} \caption{The invariant mass-squared spectrum $M_i^2$ for positronium versus the projection of the total spin, $J_z$, including annihilation with exponential, Gaussian and sharp cutoffs. The number of integration points is $N_1=N_2=21$.} \label{fig:2} \end{figure} \newpage
\section{Introduction} A {\em Gerstenhaber algebra} is a triple $({\cal A}=\oplus_{k\in{\bf Z}} {\cal A}^{k},\wedge,[\;,\;])$ where $\wedge$ is an associative, graded commutative algebra structure (e.g., over {\bf R}), $[\;,\;]$ is a graded Lie algebra structure for the {\em shifted grades} $[k]:=k+1$ (the sign $:=$ denotes a definition), and $$[a,b\wedge c]=[a,b]\wedge c+(-1)^{kj}b\wedge[a,c], \leqno{(1.1)}$$ $\forall a\in{\cal A}^{k+1}$, $b\in {\cal A}^{j}$, $c\in{\cal A}$. If this structure is supplemented by an endomorphism $\delta:{\cal A} \rightarrow{\cal A}$, of grade $-1$, such that $\delta^{2}=0$ and $$[a,b]=(-1)^{k}(\delta(a\wedge b)-\delta a\wedge b-(-1)^{k}a\wedge\delta b)\hspace{3mm}(a\in{\cal A}^{k},\,b\in{\cal A}),\leqno{(1.2)}$$ one gets an {\em exact Gerstenhaber algebra} or {\em Batalin-Vilkovisky algebra} ({\em BV-algebra}) with the {\em exact generator} $\delta$. If we also have a differential $d:{\cal A}^{k}\rightarrow {\cal A}^{k+1}$ ($d^{2}=0)$ such that $$d(a\wedge b)=(da)\wedge b+(-1)^{k}a\wedge(db) \hspace{3mm} (a\in{\cal A}^{k},\,b\in{\cal A}),\leqno{(1.3)}$$ we will say that we have a {\em differential BV-algebra}. (Some authors include $d$ in the BV-structure \cite{YK}, \cite{Xu}.) Finally,if $$d[a,b]=[da,b]+(-1)^{k}[a,db]\hspace{3mm}(a\in{\cal A}^{k},b\in{\cal A})\leqno{(1.4)}$$ the differential BV-algebra is said to be strong \cite{Xu}. On the other hand, a Jacobi manifold (e.g., \cite{GL}) is a smooth manifold $M^{m}$ (everything is of class $C^{\infty}$ in this paper) with a Lie algebra structure of local type on the space of functions $C^{\infty}(M)$ or, equivalently \cite{GL}, with a bivector field $\Lambda$ and a vector field $E$ such that $$[\Lambda,\Lambda]=2E\wedge\Lambda,\;\;[\Lambda,E]=0.\leqno{(1.5)}$$ In (1.5) one has usual Schouten-Nijenhuis brackets. If $E=0$, $(M,\Lambda)$ is a {\em Poisson manifold}. One of the most interesting examples of a BV-algebra is that of the Gerstenhaber algebra of the cotangent Lie algebroid of a Poisson manifold, described by Koszul \cite{Kz} and Y. Kosmann-Schwarzbach \cite{YK}. More generally, Xu \cite{Xu} extends a result of Koszul \cite{Kz} and proves that the exact generators of the Gerstenhaber algebra of a Lie algebroid $A\rightarrow M$ are provided by flat connections on $\wedge^{r}A$ ($r=rank\,A$), and Huebschmann \cite{H} proves a corresponding result for {\em Lie-Rinehart algebras}. The main aim of this note is to show that a Jacobi manifold also has a canonically associated, differential, BV-algebra (which, however, is not strong) namely, the Gerstenhaber algebra of the $1$-jet Lie algebroid defined by Kerbrat-Benhammadi \cite{KB}. Then, we apply results of Xu \cite{Xu}, and Evens-Lu-Weinstein \cite{ELW} to discuss duality between the homology of this BV-algebra and the cohomology of the Lie algebroid. (The homology was also independently introduced and studied by de Le\'on, Marrero and Padron \cite{LMP0}.) In the final section, we come back to a Poisson manifold $M$ with the Poisson bivector $Q$, and show that the infinitesimal automorphisms $E$ of $Q$ yield natural Poisson bivectors of the Lie algebroid $TM\oplus{\bf R}$. These bivectors lead to triangular Lie bialgebroids and BV-algebras in the usual way \cite{YK}, \cite{Xu} Notice that BV-algebras play an important role in some recent researches of theoretical physics (e.g., \cite{G}). {\it Acknowledgements}. The final version of this paper was written during the author's visit at the Centre de Math\'ematiques, \'Ecole Polytechnique, Palaiseau, France, and he wishes to thank his host institution for invitation and support. The author is grateful to Y. Kosmann-Schwarzbach for her invitation to \'Ecole Polytechnique, and several useful discussions on the modular class of a triangular Lie algebroid, as well as to J. C. Marrero for the comparison of the BV-homology and that of \cite{LMP0}, and to J. Monterde for his careful reading of the final text and his remarks. \section{The Jacobi BV-algebra} For any Lie algebroid $A\rightarrow M$ with anchor $\alpha:A\rightarrow TM$ one has a Gerstenhaber algebra ${\cal A}(A)$ defined by $${\cal A}(A):=(\oplus_{k\in{\bf N}}\Gamma\wedge^{k}A,\wedge,[\;,\;]_{SN}), \leqno{(2.1)}$$ where $\Gamma$ denotes spaces of global cross-sections, and SN denotes the Schouten-Nijenhuis bracket (e.g., \cite{YK}, \cite{Xu}; on the other hand, we refer the reader to \cite{{Mk},{YK},{ELW}}, for instance, for the basics of Lie algebroids and Lie algebroid calculus). The BV-algebra which we want to discuss is associated with the $1$-jet Lie algebroid of a Jacobi manifold $(M,\Lambda,E)$ defined in \cite{KB}, which we present as follows. We identify $M$ with $M\times\{0\}\subseteq M\times{\bf R}$, where $M\times{\bf R}$ is endowed with the Poisson bivector \cite{GL} $$P:=e^{-t}(\Lambda+\frac{\partial}{\partial t}\wedge E) \hspace{3mm}(t\in{\bf R}).\leqno{(2.2)}$$ Let $J^{1}M=T^{*}M\oplus{\bf R}$ be the vector bundle of $1$-jets of real functions on $M$, and notice that $\Gamma J^{1}M$ is isomorphic as a $C^{\infty}(M)$-module with $$\Gamma_{0}(M):=\{e^{t}(\alpha+fdt)\,/\alpha\in\wedge^{1}M,\,f\in C^{\infty}(M)\}\subseteq\wedge^{1}(M\times{\bf R}). \leqno{(2.3)}$$ A straightforward computation shows that $\Gamma_{0}(M)$ is closed under the bracket of the cotangent Lie algebroid of $(M\times{\bf R},P)$ (e.g., \cite{V3}) namely, $$\{e^{t}(\alpha+fdt),e^{t}(\beta+gdt)\}_{P}= e^{t}[L_{\sharp_{\Lambda}\alpha}\beta-L_{\sharp_{\Lambda}\beta}\alpha -d(\Lambda(\alpha,\beta)) \leqno{(2.4)}$$ $$+fL_{E}\beta-gL_{E}\alpha-\alpha(E)\beta+\beta(E)\alpha +(\{f,g\}-\Lambda(df-\alpha,dg-\beta))dt],$$ where $<\sharp_{\Lambda}\alpha,\beta>:=\Lambda(\alpha,\beta)$ $(\alpha,\beta\in\wedge^{1}M)$, and $$\{f,g\}=\Lambda(df,dg)+f(Eg)-g(Ef)\hspace{3mm}(f,g\in C^{\infty}(M))$$ is the bracket which defines the Jacobi structure \cite{GL}. Therefore, (2.4) produces a Lie bracket on $\Gamma J^{1}M$. Moreover, if $\sharp_{P}$ is defined similar to $\sharp_{\Lambda}$, we get $$\sharp_{P}(e^{t}(\alpha+fdt))=\sharp_{\Lambda}\alpha+fE-\alpha(E) \frac{\partial}{\partial t},\leqno{(2.5)}$$ and $$\rho:=(pr_{TM}\circ\sharp_{P})_{t=0}:J^{1}M\rightarrow TM \leqno{(2.6)}$$ has the properties of an anchor, since so does $\sharp_{P}$. Formulas (2.4), (2.6) precisely yield the Lie algebroid structure defined in \cite{KB}. In what follows we refer to it as the {\em $1$-jet Lie algebroid}. The mapping $f\mapsto e^{t}(df+fdt)$ is a Lie algebra homomorphism from the Jacobi algebra of $M$ to $\Gamma_{0}M$. \proclaim 2.1 Proposition. The Gerstenhaber algebra ${\cal A}(J^{1}M)$ is isomorphic to the subalgebra ${\cal A}_{0}(M):= \oplus_{k\in {\bf N}}\wedge^{k}\Gamma_{0}(M)$ of the Gerstenhaber algebra ${\cal A}(T^{*}(M\times{\bf R}))$.\par \noindent{\bf Proof.} The elements of ${\cal A}^{k}_{0}(M)$ are of the form $$\lambda=e^{kt}(\lambda_{1}+\lambda_{2}\wedge dt)\hspace{3mm} (\lambda_{1}\in\wedge^{k}M,\,\lambda_{2}\in\wedge^{k-1}M), \leqno{(2.7)}$$ and we see that ${\cal A}_{0}(M)$ is closed by the wedge product and by the bracket $\{\;,\;\}_{P}$ of general differential forms on the Poisson manifold $(M\times{\bf R},P)$ (e.g., \cite{V3}). Accordingly, $({\cal A}(J^{1}M),\wedge,\{\:,\:\})$ and $({\cal A}_{0}(M),\wedge,\{\:,\:\}_{P})$ are isomorphic Gerstenhaber algebras since they are isomorphic at the grade $1$ level, and the brackets of terms of higher degree are spanned by those of degree $1$. Q.e.d. \proclaim 2.2 Remark. Since ${\cal A}_{0}(M)$ is a Gerstenhaber algebra, the pair $({\cal A}_{0}^{0}=C^{\infty}(M),\,{\cal A}_{0}^{1}=\Gamma_{0}(M))$ is a Lie-Rinehart algebra \cite{H}. \par Now, we can prove \proclaim 2.3 Proposition. The Gerstenhaber algebra ${\cal A}_{0}(M)$ has a canonical exact generator. \par \noindent{\bf Proof.} It is known that ${\cal A}(T^{*}(M\times {\bf R}))$ has the exact generator of Koszul and Brylinski (e.g., \cite{V3}) $$\delta_{P}=i(P)d-di(P),\leqno{(2.8)}$$ where $P$ is the bivector (2.2). Hence, all we have to do is check that $\delta_{P}\lambda\in{\cal A}_{0}^{k-1}(M)$ if $\lambda$ is given by (2.7). First, we notice that $$i(P)(dt\wedge\mu)=e^{-t}(i(E)\mu+dt\wedge(i(\Lambda)\mu)) \hspace{3mm}(\mu\in\Lambda^{*}M).\leqno{(2.9)}$$ Then, if we also introduce the operator $\delta_{\Lambda}:=i(\Lambda)d-di(\Lambda)$ \cite{CLM}, and compute for $\lambda$ of (2.7), we get $$\delta_{P}\lambda=e^{(k-1)t}[\delta_{\Lambda}\lambda_{1}+(-1)^{k}L_{E} \lambda_{2}+ki(E)\lambda_{1} \leqno{(2.10)}$$ $$+(\delta_{\Lambda}\lambda_{2}+(-1)^{k}i(\Lambda)\lambda_{1}+ (k-1)i(E)\lambda_{2})\wedge dt].$$ Q.e.d. Of course, $\delta_{P}$ of (2.10) translates to an exact generator $\delta$ of the Gerstenhaber algebra ${\cal A}(J^{1}M)$, and the latter becomes a BV-algebra. This is the BV-algebra announced in Section 1, and we call it the {\em Jacobi BV-algebra of the Jacobi manifold} $(M,\Lambda,E)$. We can look at it under the two isomorphic forms indicated by Proposition 2.1. It is easy to see that the Jacobi BV-algebra above has the differential $$\bar d\lambda:=e^{(k+1)t}d(e^{-kt}\lambda), \leqno{(2.11)}$$ where $\lambda$ is given by (2.7). But, $\bar d$ is not a derivation of the Lie bracket $\{\;,\;\}$ of ${\cal A}(J^{1}M)$, and computations lead to $$(\delta_{P}\bar d+\bar d\delta_{P})\lambda=e^{kt}[(k+1)i(E)d\lambda_{1} \leqno{(2.12)}$$ $$+(L_{E}\lambda_{2}+(k+1)i(E) d\lambda_{2}-(-1)^{k}\delta_{\Lambda}\lambda_{1}) \wedge dt],$$ where $\lambda$ is given by (2.7) again. \proclaim 2.4 Remark. If we refer to the Poisson case $E=0$, we see that both $T^{*}M$ and $J^{1}M$ have natural structures of Lie algebroids. The Lie bracket and anchor map of $J^{1}M$ are given by $$\{e^{t}(\alpha+fdt),e^{t}(\beta+gdt)\}=e^{t}[\{\alpha,\beta\}_{\Lambda} +((\sharp_{\Lambda}\alpha)g-(\sharp_{\Lambda}\beta)f-\Lambda(\alpha,\beta) )dt],\leqno{(2.13)}$$ $$\rho(e^{t}(\alpha+fdt))=\sharp_{\Lambda}\alpha, \leqno{(2.14)}$$ and the mapping $\alpha\mapsto e^{t}(\alpha+0dt)$ preserves the Lie bracket, hence, $T^{*}M$ is a Lie subalgebra of $J^{1}M$, and the latter is an extension of the former by the trivial line bundle $M\times{\bf R}$. $J^{1}M$ was not yet used in Poisson geometry. \par \section{The homology of the Jacobi BV-algebra} We call the homology of the Jacobi BV-algebra of a Jacobi manifold $(M,\Lambda,E)$, with boundary operator $\delta$, {\em Jacobi homology} $H^{J}_{k}(M,\Lambda,E)$. (Another Jacobi homology was studied in \cite{CLM}.) Here, we look at this homology from the point of view of \cite{Xu} and \cite{ELW}, and discuss duality between the Jacobi homology and the Lie algebroid cohomology of $J^{1}M$, called {\em Jacobi cohomology}. Jacobi cohomology coincides with the one studied by de Le\'on, Marrero and Padr\'on in \cite{LMP}. If $C\in\Gamma\wedge^{k}(J^{1}M)^{*}$ is seen as a $k$-multilinear skew symmetric form on arguments (2.7) of degree $1$, at $t=0$, it may be written as $$C=\tilde C_{/t=0}:=e^{-kt}[(C_{1}+\frac{\partial}{\partial t}\wedge C_{2})]_{t=0} \hspace{3mm}(C_{1}\in{\cal V}^{k}M,\,C_{2}\in{\cal V}^{k-1}M), \leqno{(3.1)}$$ where ${\cal V}^{k}M$ denotes the space of $k$-vector fields on $M$. Furthermore, the coboundary, say $\sigma$, is given by the usual formula $$(\sigma C)(s_{0},...,s_{k})=\sum_{i=0}^{k}(-1)^{i}(\rho s_{i}) C(s_{0},...,\hat s_{i},...,s_{k}) \leqno{(3.2)}$$ $$+ \sum_{i<j=0}^{k}(-1)^{i+j}C(\{s_{i},s_{j}\},s_{0},..., \hat s_{i},...,\hat s_{j},...,s_{k}),$$ where $\rho$ is given by (2.6), and $s_{i}\in\Gamma J^{1}M$. Again, if we see the arguments as forms (2.7) with $k=1$, (3.2) becomes $$(\sigma C)=[\sigma_{P}\tilde C)]_{t=0}=[P,\tilde C]_{t=0}, \leqno{(3.3)}$$ where $\sigma_{P}$ is the Lichnerowicz coboundary (e.g., \cite{V3}). Up to the sign, (3.3) is the coboundary defined in \cite{LMP} namely, $$\sigma C=[\Lambda,C_{1}]-kE\wedge C_{1}-\Lambda\wedge C_{2} \leqno{(3.4)}$$ $$-\frac{\partial}{\partial t} \wedge([\Lambda,C_{2}]-(k-1)E\wedge C_{2} +[E,C_{1}]).$$ We denote the Jacobi cohomology spaces by $H^{k}_{J}(M,\Lambda,E)$. \proclaim 3.1 Remark. {\rm \cite{LMP}}. The anchor $\rho$ induces homomorphisms $\rho^{\sharp}:H^{k}_{de\,R}(M)\rightarrow H^{k}_{J}(M,\Lambda,E)$ given by $$(\rho^{\sharp}\lambda)(s_{1},...,s_{k})=(-1)^{k}\lambda(\rho s_{1},...,\rho s_{k}) \hspace{3mm}(\lambda\in\wedge^{k}M,\, s_{i}\in\Gamma J^{1}M). \leqno{(3.5)}$$\par Now, we need a recapitulation of several results of \cite{Xu} and \cite{ELW}. For a Lie algebroid $A\rightarrow M$ with anchor $a$, an $A$-{\em connection} $\nabla$ on a vector bundle $E\rightarrow M$ consists of {\em derivatives} $\nabla_{s}e\in\Gamma E$ ($s\in \Gamma A,\;e\in\Gamma E$) which are ${\bf R}$-bilinear and satisfy the conditions $$\nabla_{fs}e=f\nabla_{s}e,\;\; \nabla_{s}(fe)=(a(s)f)e+f\nabla_{s}e\hspace{3mm}(f\in C^{\infty}(M)).$$ For an $A$-connection, curvature may be defined as for usual connections. Any flat $A$-connection $\nabla$ on $\wedge^{r}A$ ($r=rank\,A$) produces a {\em Koszul operator} $D:\Gamma\wedge^{k}A \rightarrow\Gamma\wedge^{k-1}A$, locally given by $$DU=(-1)^{r-k+1}[i(d\omega)\Omega+\sum_{h=1}^{r}\alpha^{h} \wedge(i(\omega)\nabla_{s_{h}}\Omega],$$ where $\Omega\in\Gamma\wedge^{r}A$, $\omega\in\Gamma\wedge^{r-k}A^{*}$ is such that $i(\omega)\Omega=U$, $s_{h}$ is a local basis of $A$, and $\alpha^{h}$ is the dual cobasis of $A^{*}$. Moreover, $D$ is an exact generator of the Gerstenhaber algebra of $A$, and every exact generator is defined by a flat $A$-connection as above. The operator $D$ is a boundary, and yields a corresponding homology, called the {\em homology of the Lie algebroid $A$ with respect to the flat $A$-connection} $\nabla$, $H_{k}(A,\nabla)$. For two flat connections $\nabla,\,\bar\nabla$ such that $D-\bar D=i(\alpha)$, where $\alpha=d_{A}f$ $(f\in C^{\infty}(M) )$, one has $H_{k}(M,\nabla)=H_{k}(M,\bar\nabla)$. If $\exists\Omega\in\Gamma \wedge^{r}A$ which is nowhere zero, and $\nabla\Omega=0$, one has the duality $H_{k}(A,\nabla)=H^{r-k}(A)$, defined by sending $Q\in\Gamma\wedge^{k}A$ to $*_{\Omega}Q:=i(Q)\Omega$. These results may be applied to the cotangent Lie algebroid of an orientable Poisson manifold $(N^{n},Q)$. In this case, the flat connection $\nabla_{\theta}\Psi=\theta\wedge(di(Q)\Psi)$ ($\theta\in T^{*}N$, $\Psi\in\wedge^{n}N$) precisely has the Koszul operator $\delta_{Q}$ and defines the known Poisson homology $H_{k}(N,Q)$ (e.g., \cite{V3}). Finally (\cite{Xu}, Proposition 4.6 and Theorem 4.7), if $N$ has the volume form $\Omega$, which defines a connection $\nabla_{0}$ by $\nabla_{0}\Omega=0$, and if $W^{Q}$ is the {\em modular vector field} which acts on $f\in C^{\infty}(M)$ according to the equation $L_{X^{Q}_{f}}\Omega=(W^{Q}f)\Omega$ ($X^{Q}_{f}$ is the Hamiltonian field of $f$) \cite{W}, one has $\delta_{Q}-D_{0}=i(W^{Q})$. Accordingly, if the modular field $W^{Q}$ is Hamiltonian (i.e., $(N,Q)$ is a {\em unimodular Poisson manifold}), $H_{k}(N,Q)=H^{n-k}(T^{*}N)$. The case of a general, possibly non orientable, Poisson manifold is studied in \cite{ELW}. The expression of $\nabla_{\theta}\Psi$ above can be seen as the local equation of a connection on $\wedge^{n}T^{*}N$, and it still defines the Koszul operator $\delta_{Q}$. The general duality Theorem 4.5 of \cite{ELW} is $$H_{k}(N,Q)=H^{n-k}(T^{*}N,\wedge^{n}T^{*}N),\leqno{(*)}$$ where the right hand side is the cohomology of the Lie algebroid $T^{*}N$ with values in the line bundle $\wedge^{n}T^{*}N$. This means that the $k$-cocycles are spanned by cross sections $V\otimes\Psi$, $V\in {\cal V}^{k}N$, $\Psi\in\Gamma\wedge^{n}T^{*}N$, and the coboundary is given by $$\partial(V\otimes\Psi)=[Q,V]\otimes\Psi+(-1)^{k}V\otimes\nabla\Psi.$$ Duality ($*$) is again defined by the isomorphism which sends $V\otimes\Psi$ to $i(V)\Psi$. With this recapitulation finished, we apply the results to Jacobi manifolds $(M^{m},\Lambda,E)$. Consider the Poisson manifold $(M\times{\bf R},P)$ which we already used before. Then $\delta_{P}$ is the Koszul operator of the $(T^{*}M\times{\bf R})$-connection $$\nabla_{\theta}\Psi=\theta\wedge(di(P)\Psi)\hspace{3mm}(\theta\in T^{*} (M\times{\bf R}),\,\Psi\in\wedge^{m+1}(M\times{\bf R})).\leqno{(3.6)}$$ In particular, if we take $$\theta=e^{t}(\alpha+fdt),\hspace{2mm}\Psi=e^{(m+1)t}\Phi\wedge dt \hspace{3mm}(\alpha\in T^{*}M,\,\Phi\in\wedge^{m}M)\leqno{(3.7)}$$ (and use (2.9)) we get $$\nabla_{\theta}\Psi=e^{(m+1)t}[fdi(E)\Phi-\alpha\wedge(di(\Lambda)\Phi +mi(E)\Phi)]\wedge dt,\leqno{(3.8)}$$ and this formula may be seen as defining a $J^{1}M$-connection on $\wedge^{m+1}J^{1}M$. Clearly, the Koszul operator of this connection must be $\delta$ of (2.10). Therefore, we have \proclaim 3.2 Proposition. The Jacobi homology of $(M,\Lambda,E)$ is equal to the homology of the Lie algebroid $J^{1}M$ with respect to the flat connection (3.8) i.e., $$H^{J}_{k}(M,\Lambda,E)=H_{k}(J^{1}M,\nabla).\leqno{(3.9)}$$ \par Now, assume that $M$ has a volume form $\Phi\in\wedge^{m}M$. Then $\Omega:=e^{(m+1)t}\Phi\wedge dt$ is a volume form on $M\times{\bf R}$, and one has a connection $\nabla_{0}$ defined by $\nabla_{0}\Omega=0$ with a Koszul operator $D_{0}$ such that $$\delta_{P}-D_{0}=i(W^{P}),\leqno{(3.10)}$$ where $W^{P}$ is the corresponding modular vector field i.e., $$L_{X^{P}_{\varphi}}\Omega=(W^{P}\varphi)\Omega \hspace{3mm}(\varphi\in C^{\infty}(M\times{\bf R})).\leqno{(3.11)}$$ We need the interpretation of (3.10) at $t=0$. To get it, we take local coordinates $(x^{i})$ on $M$, and compute the local components of $W^{P}$ by using (3.11) for $\varphi=x^{i}$ and $\varphi=t$. Generally, we have $$X^{P}_{\varphi}=i(d\varphi)P=e^{-t}(\sharp_{\Lambda}d\varphi+ \frac{\partial\varphi}{\partial t}E-(E\varphi)\frac{\partial}{\partial t}). \leqno{(3.12)}$$ On the other hand, on $M$, let us define a vector field $V$ and a function $div_{\Phi}E$ by $$L_{\sharp_{\Lambda}df}\Phi=(Vf)\Phi,\;L_{E}\Phi=(div_{\Phi}E)\Phi \hspace{3mm}(f\in C^{\infty}(M). \leqno{(3.13)}$$ (The fact that $V$ is a derivation of $C^{\infty}(M)$ follows easily from the skew symmetry of $\Lambda$.) Then, the calculation of the local components of $W^{P}$ yield $$W^{P}=e^{-t}[V-mE+(div_{\Phi}E)\frac{\partial}{\partial t}].\leqno{(3.14)}$$ At $t=0$, (3.14) defines a section of $TM\oplus{\bf R}$ which we denote by $V^{(\Lambda,E)}$, and call the {\em modular field} (not a vector field, of course) of the Jacobi manifold. As in the Poisson case, if $\Phi\mapsto a\Phi$ ($a>0$), $V^{(\Lambda,E)}\mapsto V^{(\Lambda,E)}+\sigma(\ln a)$ hence, what is well defined is the Jacobi cohomology class $[V^{(\Lambda,E)}]$, to be called the {\em modular class}. If the modular class is zero $(M,\Lambda,E)$ is a {\em unimodular Jacobi manifold}. It is also possible to get the modular class $[V^{(\Lambda,E)}]$ from the general definition of the modular class of a Lie algebroid \cite{ELW}. In the case of the algebroid $J^{1}M$, the definition of \cite{ELW} means computing the expression $${\cal E}:=(L^{J^{1}M}_{e^{t}(df+fdt)}[(e^{mt}\Phi) \wedge(e^{t}dt)])\otimes\Phi$$ $$+(e^{(m+1)t}\Phi\wedge dt)\otimes(L_{\rho(e^{t}(df+fdt))}\Phi,$$ where $\rho$ is given by (2.6), and $$L^{J^{1}M}_{e^{t}(df+fdt)}[(e^{mt}\Phi) \wedge(e^{t}dt)]=\{e^{t}(df+fdt),(e^{mt}\Phi) \wedge(e^{t}dt)\}_{P}.$$ If we decompose $(e^{mt}\Phi)=\wedge^{n}_{i=1}(e^{t}\varphi_{i})$, $\varphi_{i}\in\wedge^{1}M$, the result of the required computation turns out to be $${\cal E}=(2(Vf)+2f(div_{\Phi}E)-Ef)(e^{(m+1)t}\Phi\wedge dt) \otimes\Phi.$$ By comparing with (3.14), we see that the modular class in the sense of \cite{ELW} is the Jacobi cohomology class of the cross section of $TM\oplus {\bf R}$ defined by $$A^{(\Lambda,E)}=2V^{(\Lambda,E)}-(2m+1)E.$$ With all this notation in place, the recalled results of \cite{Xu}, Proposition 4.6, and \cite{ELW}, Theorem 4.5, yield \proclaim 3.3 Proposition. If $(M,\Lambda,E)$ is a unimodular Jacobi manifold one has duality between Jacobi homology and cohomology: $$H^{J}_{k}(M,\Lambda,E)=H^{m-k+1}_{J}(M,\Lambda,E).\leqno{(3.15)}$$ If $(M,\Lambda,E)$ is an arbitrary Jacobi manifold, one has the duality $$H^{J}_{k}(M,\Lambda,E)=H^{m-k+1}_{J}(J^{1}M, \wedge^{m+1}J^{1}M).\leqno{(3.15')}$$ \par \noindent{\bf Proof.} The right hand side of (3.15$'$) is {\em Jacobi cohomology with values in} $\wedge^{m+1}J^{1}M$, similar to that in ($*$). The homologies and cohomologies of (3.15) and (3.15$'$) are to be seen as given by subcomplexes of $\oplus_{k}\wedge^{k}(M\times{\bf R})$, $\oplus_{k}{\cal V}^{k}(M\times{\bf R})$ defined by (2.7) and (3.1). Then the result follows by the proofs of the theorems of \cite{Xu}, \cite{ELW} quoted earlier, if we notice that $$i[e^{-kt}(C_{1}+\frac{\partial}{\partial t}\wedge C_{2})] (e^{(m+1)t}\Phi\wedge dt)$$ $$=e^{(m-k+1)t}[(-1)^{m}i(C_{2})\Phi+(i(C_{1})\Phi)\wedge dt].$$ The notation is that of (3.1) and (3.7). Q.e.d. To get examples, let us look at the {\em transitive Jacobi manifolds} \cite{GL}. a). Let $M^{2n}$ be a locally or globally conformal symplectic manifold with the $2$-form $\Omega=e^{\sigma_{\alpha}}\Omega_{\alpha}$, where $\Omega_{\alpha}$ are symplectic forms on the sets $U_{\alpha}$ of an open covering of $M$, and $\sigma_{\alpha}\in C^{\infty}(U_{\alpha})$. Then (e.g., \cite{V2}) $\{d\sigma_{\alpha}\}$ glue up to a global closed $1$-form $\omega$, which is exact iff $\exists\alpha$, $U_{\alpha}=M$, and $\sharp_{\Lambda}:=\flat_{\Omega}^{-1}$, $E:=\sharp_{\Lambda}\omega$ define a Jacobi structure on $M$ \cite{GL}. It follows easily that $L_{E}\Omega=0$ hence, $div_{\Omega^{n}}E=0$. Furthermore, $$L_{\sharp_{\Lambda}df}\Omega^{n}=-n(n-1)df\wedge\omega\wedge\Omega^{n-1} .$$ Using the {\em Lepage decomposition theorem} (\cite{LM}, pg.46) we see that $df\wedge\omega=\xi+\varphi\Omega$, where $$\xi\wedge\Omega^{n-1}=0,\;\;\varphi=-\frac{1}{n}i(\Lambda)(df\wedge\omega) =Ef.$$ Hence, $V=-n(n-1)E$, and $V^{(\Lambda,E)}=-n(2n-1)E$. Then, for $f\in C^{\infty}(M)$, (3.4) yields $\sigma f=\sharp_{\Lambda}df-(Ef)(\partial/\partial t)$, and $\sigma f=E$ holds iff $\omega=df$. Thus, (3.15) holds on globally conformal symplectic manifolds. But, (3.15) may not hold in the true locally conformal symplectic case. For instance, it follows from Corollary 3.15 of \cite{LMP0} that the result does not hold on a Hopf manifold with its natural locally conformal K\"ahler structure. (Private correspondence from J. C. Marrero.) b). Let $M^{2n+1}$ be a contact manifold with the contact $1$-form $\theta$ such that $\Phi:=\theta\wedge(d\theta)^{n}$ is nowhere zero. Then $M$ has the Reeb vector field $E$ where $$i(E)\theta=1,\;\;i(E)d\theta=0,$$ and $\forall f\in C^{\infty}(M)$ there is a {\em Hamiltonian vector field} $X^{\theta}_{f}$ such that $$i(X^{\theta}_{f})\theta=f,\;\;i(X^{\theta}_{f})d\theta=-df+(Ef)\theta.$$ Furthermore, if $$\Lambda(df,dg):=d\theta(X^{\theta}_{f},X^{\theta}_{g}) \hspace{3mm}(f,g\in C^{\infty}(M)),$$ $(\Lambda,E)$ is a Jacobi structure \cite{GL}. Now, let $(q^{i},p_{i},z)$ $(i=1,...,n)$ be local canonical coordinates, such that $\theta=dz-\sum_{i}p_{i}dq^{i}$. Then $$E=\frac{\partial}{\partial z},\;\Lambda=\sum_{i}\frac{\partial}{\partial q^{i}}\wedge\frac{\partial}{\partial p_{i}}+\frac{\partial}{\partial z} \wedge(\sum_{i}p_{i}\frac{\partial}{\partial p{_i}}).$$ This leads to $div_{\Phi}E=0$, $V^{(\Lambda,E)}=nE$, and it follows that there is no $f\in C^{\infty}(M)$ satisfying $\sigma f=(nE\oplus 0)$. We close this section by the remark that the identification of a manifold $M$ with $M\times\{0\}\subseteq M\times {\bf R}$ leads to other interesting structures too. For instance, if we define the spaces $$\wedge_{0}^{k}M:=\{e^{t}(\xi_{1}+\xi_{2}\wedge dt) \,/\,\xi_{1}\in\wedge^{k}M,\xi_{2}\in\wedge^{k-1}M\},$$ the triple $(\oplus_{k} \wedge_{0}^{k}M,d,i(X+f(\partial/\partial t))$ is a {\em Gelfand-Dorfman complex} \cite{Dor}, and a Jacobi structure on $M$ is equivalent with a {\em Hamiltonian structure} \cite{Dor} on this complex. On the other hand, if we have a Jacobi manifold $(M,\Lambda,E)$, and put $${\cal V}_{0}^{k}M:=\{e^{-(k-1)t}(Q_{1}+\frac{\partial}{\partial t} \wedge Q_{2})\,/\,Q_{1}\in{\cal V}^{k}M,\,Q_{2}\in{\cal V}^{k-1}M\},$$ then $(\oplus_{k}{\cal V}_{0}^{k}M,[\:,\:],\sigma_{P})$ ($P$ is defined by (2.2)) is a differential graded Lie algebra, the cohomology of which is exactly the $1$-{\em differentiable Chevalley-Eilenberg cohomology} $H^{k}_{1-dif}(M,\Lambda,E)$ of Lichnerowicz \cite{Lz}. In particular, $H^{1}_{1-dif}(M,\Lambda,E)$ is the quotient of the space of conformal Jacobi infinitesimal automorphisms by the space of the Jacobi Hamiltonian vector fields \cite{Lz}. \section{Lie bialgebroid structures on TM$\oplus${\bf R}} In the Poisson case, $T^{*}M$ is a Lie bialgebroid \cite{YK}, \cite{MX} with dual $TM$. This is not true for $J^{1}M$ on Jacobi manifolds in spite of the fact that $(J^{1}M)^{*}=TM\oplus{\bf R}$ has a natural Lie algebroid structure, which extends the one of $TM$. Namely, if we see ${\cal X}\in\Gamma(TM\oplus{\bf R})$ as a vector field of $M\times {\bf R}$ given by $${\cal X}=(X+f\frac{\partial}{\partial t})_{t=0}\hspace{3mm} (X\in\Gamma TM,\,f\in C^{\infty}(M)),\leqno{(4.1)}$$ we have the Lie bracket $$[{\cal X},{\cal Y}]_{0}:= [X+f\frac{\partial}{\partial t},Y+g\frac{\partial}{\partial t}]= [X,Y]+(Xg-Yf)\frac{\partial}{\partial t}, \leqno{(4.2)}$$ and the anchor map $a({\cal X}):=X$. If we would be in the case of a Lie bialgebroid, the bracket $\{f,g\}_{s}$ $:=<df,d_{*}g>$ $(f,g\in C^{\infty}(M))$, where $d,\,d_{*}$ are the differentials of the Lie algebroids $TM\oplus{\bf R}$ and $J^{1}M$, respectively, would be Poisson \cite{YK}, \cite{MX}. This is not true since one gets $\{f,g\}_{s}=\Lambda(df,dg)$. \proclaim 4.1 Remark. The differential $\bar d$ defined by (2.11) is the same as the differential $d$ of the Lie algebroid $TM\oplus{\bf R}$ with the bracket (4.2).\par In Poisson geometry, the cotangent Lie bialgebroid structure is produced by a Poisson bivector $\Pi$ of $TM$ i.e., $[\Pi,\Pi]=0$. It is natural to ask what is the structure produced by a Poisson bivector $\Pi$ of $TM\oplus{\bf R}$. We generalize this question a bit namely, we fix a closed $2$-form $\Omega$ on $M$, and take the Lie bracket $$[{\cal X},{\cal Y}]_{\Omega}:=[{\cal X},{\cal Y}]_{0}+\Omega(X,Y) \frac{\partial}{\partial t}. \leqno{(4.2')}$$ The notation, and the anchor map $a$ are the same as for (4.2). It is known that (4.2$'$) defines all the transitive Lie algebroid structures over $M$ such that the kernel of the anchor is a trivial line bundle, up to an isomorphism \cite{Mk}. A Poisson bivector $\Pi$ on $TM\oplus{\bf R}$ with the bracket (4.2$'$) will be called an $\Omega$-{\em Poisson structure} on $M$. \proclaim 4.2 Proposition. An $\Omega$-Poisson structure $\Pi$ on $M$ is equivalent with a pair $(Q,E)$, where $Q$ is a Poisson bivector on $M$ (i.e., $[Q,Q]=0$), and $E$ is a vector field such that $$L_{E}Q=\sharp_{Q}\Omega. \leqno{(4.3)}$$ \par \noindent{\bf Proof.} Using the identification (4.1) of the cross sections of $TM\oplus{\bf R}$ with vector fields on $M\times{\bf R}$ for $t=0$, and local coordinates $(x^{i})$ on $M$, we may write $$\Pi=Q+\frac{\partial}{\partial t}\wedge E= \frac{1}{2}Q^{ij}(x)\frac{\partial}{\partial x^{i}}\wedge \frac{\partial}{\partial x^{j}}+\frac{\partial}{\partial t} \wedge(E^{k}(x)\frac{\partial}{\partial x^{k}}),\leqno{(4.4)}$$ where $Q$ is a bivector field on $M$, $E$ is a vector field, and the Einstein summation convention is used. Now, $[\Pi,\Pi]_{\Omega}=0$ can be expressed by the known formula of the Schouten-Nijenhuis bracket of decomposable multivectors (e.g., \cite{V3}, formula (1.12)), and (4.2$'$). The result is equivalent to $[Q,Q]=0$ and (4.3). Q.e.d. \proclaim 4.3 Corollary. If $(M,Q)$ is a Poisson manifold, $Q$ extends to a $\Omega$-Poisson structure for every closed $2$-form $\Omega$, where the de Rham class $[\Omega]$ has zero $\sharp_{Q}$-image in the Poisson cohomology of $(M,Q)$, by taking $E$ such that (4.3) holds. \par This is just a reformulation of Proposition 4.2. It is well known that a Poisson bivector on a Lie algebroid $A$ induces a bracket on $\Gamma A^{*}$ such that $(A,A^{*})$ is a triangular Lie bialgebroid \cite{YK}, \cite{MX}. Namely, the Poisson bivector $\Pi$ of (4.4) yields the following bracket $$\{\alpha\oplus f,\beta\oplus g\}_{\Omega}:= L^{\Omega}_{\sharp_{\Pi}(\alpha\oplus f)}(\beta\oplus g) -L^{\Omega}_{\sharp_{\Pi}(\beta\oplus g)}(\alpha\oplus f) -d_{\Omega}(\Pi(\alpha\oplus f,\beta\oplus g)),\leqno{(4.5)}$$ where $\alpha\oplus f,\beta\oplus g\in\Gamma J^{1}M$, and the index $\Omega$ denotes the fact that the operators involved are those of the Lie algebroid calculus of (4.2$'$). To make this formula explicit, notice that $$\sharp_{\Pi}(\alpha+fdt)=\sharp_{Q}\alpha+fE-\alpha(E) \frac{\partial}{\partial t},\leqno{(4.6)}$$ whence $$\Pi(\alpha+fdt,\beta+gdt)= Q(\alpha,\beta)+f\beta(E)-g\alpha(E).\leqno{(4.7)}$$ Then, by evaluation on a field of the form (4.1), and with (4.2$'$), we obtain $$L_{\sharp_{\Pi}(\alpha+fdt)}^{\Omega}(\beta+gdt)= L_{\sharp_{\Pi}(\alpha+fdt)}(\beta+gdt)-g(\flat_{\Omega}\sharp_{Q} \alpha)-fgi(E)\Omega,\leqno{(4.8)}$$ where $\flat_{\Omega}X:=i(X)\Omega$. As a consequence, (4.5) becomes $$\{\alpha\oplus f,\beta\oplus g\}_{\Omega}:= [\{\alpha,\beta\}_{Q}+f(L_{E}\beta+\flat_{\Omega}\sharp_{Q}\beta) \leqno{(4.9)}$$ $$-g(L_{E}\alpha+\flat_{\Omega}\sharp_{Q}\alpha)]\oplus [(\sharp_{Q}\alpha)g-(\sharp_{Q}\beta)f+f(Eg)-g(Ef)],$$ where $(Q,E)$ are associated with $\Pi$ as in Proposition 4.2. The anchor map of the Lie algebroid $J^{1}M$ with (4.9) is $\rho:=pr_{TM}\circ\sharp_{\Pi}$, and it is provided by (4.6). In particular, Proposition 4.2 tells us that a pair $(Q,E)$ which consists of a Poisson bivector $Q$ and an infinitesimal automorphism $E$ of $Q$, to which we will refer as an {\em enriched Poisson structure}, provides a Poisson bivector $\Pi$ on $TM\oplus{\bf R}$ with the bracket (4.2), and a Lie bialgebroid $(TM\oplus{\bf R},J^{1}M=T^{*}M\oplus{\bf R})$. An example (suggested by \cite{Lh}) can be obtained as follows. Let $(M,\Lambda,E)$ be a Jacobi manifold. A {\em time function} is a function $\tau\in C^{\infty}(M)$ which has no critical points and satisfies $E\tau=1$. If such a function exists, $(\Lambda_{0}:=\Lambda-(\sharp_{\Lambda}dt)\wedge E, E)$ is an enriched Poisson structure. Jacobi manifolds with time may be seen as generalized phase spaces of time-dependent Hamiltonian systems. Namely, if $H\in C^{\infty}(M)$ is the Hamiltonian function, the trajectories of the system are the integral lines of the vector field $X^{0}_{H}:=\sharp_{\Lambda_{0}}df+E$. Let us briefly indicate the important objects associated with the Lie algebroids $TM\oplus{\bf R}$, defined by the bracket (4.2$'$), and $J^{1}M$ with the bracket (4.9). The cohomology of $TM\oplus{\bf R}$ is that of the cochain spaces $$\wedge^{k}_{\Omega}M:=\{\lambda=\lambda_{1}+\lambda_{2}\wedge dt \;/\;\lambda_{1}\in \wedge^{k}M,\lambda_{2}\in\wedge^{k-1}M\} \leqno{(4.10)}$$ with the corresponding coboundary, say $d_{\Omega}$. A straightforward evaluation of $d_{\Omega}\lambda$ on arguments $X_{i}+f_{i}(\partial/\partial t)$, in accordance with the Lie algebroid calculus \cite{Mk}, yields the formula $$d_{\Omega}\lambda=d\lambda-(-1)^{k}\Omega\wedge\lambda_{2}. \leqno{(4.11)}$$ The {\em Poisson cohomology} of $TM\oplus{\bf R}$ above i.e., the cohomology of the Lie algebroid $J^{1}M$ with (4.9), can be seen (with (4.1)) as having the cocycle spaces $${\cal C}^{k}(M):=\{C=C_{1}+\frac{\partial}{\partial t}\wedge C_{2}\; /\;C_{1}\in{\cal V}^{k}M,C_{2}\in {\cal V}^{k-1}M\},\leqno{(4.12)}$$ and the coboundary $\partial C=[\Pi,C]_{\Omega}$, with $\Pi$ of (4.4) and the $\Omega$-Schouten-Nijenhuis bracket. In order to write down a concrete expression of this coboundary, we define an operation $U\wedge_{\Omega}V\in{\cal V}^{k+h-2}$, for $U\in {\cal V}^{k}M$, $V\in{\cal V}^{h}M$, by the formula $$U\wedge_{\Omega}V(\alpha_{1},...,\alpha_{k+h-2})=\frac{1} {(k-1)!(h-1)!}\sum_{\sigma\in S_{k+h-2}}[(sign\,\sigma)\leqno{(4.13)}$$ $$\cdot\sum_{i=1}^{m}U(\epsilon^{i}, \alpha_{\sigma_{1}},...,\alpha_{\sigma_{k-1}}) V(\flat_{\Omega}e_{i},\alpha_{\sigma_{k}},...,\alpha_{\sigma_{k+h-2}})], $$ where $S$ is the symmetric group, $e_{i}$ is a local tangent basis of $M$, and $\epsilon^{i}$ is the corresponding dual cobasis. If $U,V$ are vector fields, $U\wedge_{\Omega}V=\Omega(U,V)$. By computing for decomposable multivectors $C_{1},C_{2}$, we get $$\partial C=[Q,C_{1}]+\frac{\partial}{\partial t}\wedge ([Q,C_{2}]+Q\wedge_{\Omega} C_{1}-L_{E}C_{1}),\leqno{(4.14)}$$ where the brackets are the usual Schouten-Nijenhuis brackets on $M$. Furthermore, the exact generator of the BV-algebra of the Lie algebroid $J^{1}M$ is $\delta_{\Omega}:=[i(\Pi),d_{\Omega}]$, and using (4.11) we get $$\delta_{\Omega}(\lambda_{1}+\lambda_{2}\wedge dt)= \delta_{Q}\lambda_{1}+(-1)^{k-1}([i(Q),e(\Omega)]\lambda_{2}\leqno{(4.15)}$$ $$-di(E)\lambda_{2}) +(\delta_{Q}\lambda_{2})\wedge dt,$$ where $e(\Omega)$ is exterior product and $[\;,\;]$ is the commutant of the operators. Finally, let us discuss the {\em modular class} of the Lie algebroid ($J^{1}M$, (4.9)). For simplicity, we assume the manifold $M$ orientable, with the volume form $\Phi\in\Gamma\wedge^{m}M$. In the non orientable case, the same computations hold if $\Phi$ is replaced by a density $\Phi\in\Gamma|\wedge^{m}M|$. Again, we denote by $W^{Q}$ the modular vector field defined by $L_{\sharp_{Q}df}\Phi=(W^{Q}f)\Phi$ (see Section 3). There are two natural possibilities to define a modular class for the algebroid $J^{1}M$. One is by computing the Lie derivative: $$L_{\sharp_{\Pi}(\alpha+fdt)}(\Phi\wedge dt)= [L_{\sharp_{Q}\alpha}\Phi+fL_{E}\Phi+df\wedge i(E)\Phi] \wedge dt.\leqno{(4.16)}$$ This result is obtained if the computation is done after $\Phi$ is decomposed into a product of $m$ $1$-forms, and by using (4.6). Since $i(E)(df\wedge\Phi)=0$, the last term in (4.16) is $(Ef)\Phi$, and if we also use (3.13), (4.16) yields $$L_{\sharp_{\Pi}(df+fdt)}(\Phi\wedge dt)= (W^{Q}f+fdiv_{\Phi}E+Ef)(\Phi\wedge dt).\leqno{(4.17)}$$ Therefore, we get the {\em modular field} $$W^{\Pi}:=W^{Q}+E+(div_{\Phi}E)\frac{\partial}{\partial t}.\leqno{(4.18)}$$ If $\Phi$ is replaced by $h\Phi$ $(h\in C^{\infty}(M))$, it follows easily that the $\Pi$-Poisson cohomology class $[W^{\Pi}]$ is preserved. This will be the {\em modular class}. The second possibility is to apply the general definition of \cite{ELW}. Similar to what we had for Jacobi manifolds in Section 3, this asks us to compute the flat connection $D$ on $(\wedge^{m+1}J^{1}M) \otimes(\wedge^{m}T^{*}M)$ given by $$D_{(df+fdt)}[(\Phi\wedge dt)\otimes\Phi]= L^{J^{1}M}_{(df+fdt)}(\Phi\wedge dt)\otimes\Phi+ (\Phi\wedge dt)\otimes (L_{\rho(df+fdt)}\Phi).\leqno{(4.19)}$$ From Lie algebroid calculus, we know that $$L^{J^{1}M}_{(df+fdt)}(\Phi\wedge dt)=\{df+fdt,\Phi\wedge dt\}_{\Omega}, \leqno{(4.20)}$$ where the bracket is the Schouten-Nijenhuis extension of (4.9). If we look at a decomposition $\Phi=\varphi_{1}\wedge...\wedge \varphi_{n}$ $(\varphi_{i}\in\wedge^{1}M)$, (4.9) yields $$\{df+fdt,dt\}_{\Omega}=0,$$ $$\{df+fdt,\varphi_{i}\}_{\Omega}=L_{\rho(df+fdt)}\varphi_{i} -\varphi_{i}(E)df+f\flat_{\Omega}\sharp_{Q}\beta-(\sharp_{Q}\beta(f))dt,$$ and we get $$L^{J^{1}M}_{(df+fdt)}(\Phi\wedge dt)=[L_{(df+fdt)}\Phi-(Ef)\Phi +ftr(\flat_{\Omega}\circ\sharp_{Q})\Phi]\wedge dt.\leqno{(4.21)}$$ But, we also have $$L_{\rho(df+fdt)}\Phi=L_{\sharp_{Q}df+fE}\Phi=[W^{Q}f+fdiv_{\Phi}E +Ef]\Phi.\leqno{(4.22)}$$ By inserting (4.21), (4.22) into (4.19), we get another {\em modular field} namely, $$A_{\Omega}:=(2W^{Q}+E)\oplus(div_{\Phi}E+tr(\flat_{\Omega}\circ\sharp_{Q})) =(2W^{\Pi}-E)\oplus tr(\flat_{\Omega}\circ\sharp_{Q}).\leqno{(4.23)}$$ From the general results of \cite{ELW}, it is known that the $\Pi$-Poisson cohomology class of this field is independent of the choice of $\Phi$, and it is a {\em modular class} of $J^{1}M$. As for the modular class of $TM\oplus{\bf R}$ with the bracket (4.2$'$), it vanishes for reasons similar to those for the class of the tangent groupoid $TM$ \cite{ELW}. We finish by another interpretation of the enriched Poisson structures. If ${\cal F}$, be an arbitrary associative, commutative, real algebra, we may say that $f:M\rightarrow{\cal F}$ is differentiable if for any ${\bf R}$-linear mapping $\phi:{\cal F}\rightarrow{\bf R}$, $\phi\circ f\in C^{\infty}(M)$. Furthermore, an ${\bf R}$-linear operator $v_{x}$ which acts on germs of ${\cal F}$-valued differentiable functions at $x\in M$, and satisfies the Leibniz rule will be an ${\cal F}$-{\em tangent vector} of $M$ at $x$. Then, we have natural definitions of tangent spaces $T_{x}(M,{\cal F})$, differentiable ${\cal F}$-vector fields, etc. \cite{V1}. A bracket $\{\;,\;\}$ which makes $C^{\infty}(M,{\cal F})$ into a Poisson algebra will be called an ${\cal F}$-{\em Poisson structure} on $M$. Now, take ${\cal F}$ to be the {\em Studi algebra of parabolic dual numbers} ${\cal S}:={\bf R}[s\,/\,s^{2}=0]$, where $s$ is the generator. An ${\cal S}$-Poisson structure $\Pi$ in the above mentioned sense will be called a {\em Studi-Poisson structure}. The restriction of $\Pi$ to real valued functions is a Poisson bivector $Q$ on $M$, and the Jacobi identity shows that the Hamiltonian vector field $X^{\Pi}_{s}$ of the constant function $s$ is an infinitesimal automorphism $E$ of $Q$. Conversely, the pair $(Q,E)$ defines the Studi-Poisson bracket $$\{f_{0}+f_{1}s,g_{0}+g_{1}s\}:=\{f_{0},g_{0}\}_{Q}+f_{1}(Eg_{0})- g_{1}(Ef_{0})$$ $$+s(\{f_{0},g_{1}\}_{Q}+\{f_{1},g_{0}\}_{Q}+f_{1}(Eg_{1})-g_{1}(Ef_{1})).$$ Notice that we cannot say that $v_{x}s=0$ for all $v_{x}\in T_{x}(M,{\cal F})$ since $v_{x}$ was linear over ${\bf R}$ only.
\section{INTRODUCTION} Being the only example of a non-cuprate layered perovskite superconductor \cite{1}, Sr$_2$RuO$_4$ has attracted considerable attention despite its rather low critical temperature , $T_c \sim$1 K \cite{1}. Its normal state is characterized as an essentially two-dimensional Fermi liquid and the coherent interlayer transport settles in at low temperature only \cite{2}. The susceptibility is most likely dominated by an enhanced Pauli spin susceptibility, $\tilde{\chi}$. Meanwhile, the Sommerfeld coefficient, $\gamma$, in the specific heat is enhanced by a factor 3.5 with respect to band structure calculations \cite{2,3}. It yields a Wilson ratio, $R_W$ ($\sim \tilde{\chi}/\gamma$), of 1.7. This value indicates that the enhancements in both susceptibility and electronic specific heat can be ascribed to the same origin: most likely correlations among electrons \cite{2}. Noticing that SrRuO$_3$ is ferromagnetic (FM), it has been conjectured that Sr$_2$RuO$_4$ is close to a FM instability as well \cite{4}. This assertion is corroborated by microscopic calculation of magnetic properties of ruthenates \cite{5}. Since FM fluctuations disfavor both s- and d-wave superconductivity, it has been suggested that superconductivity in Sr$_2$RuO$_4$ should possess p-wave symmetry (triplet pairing) \cite{4,6}. Conventional local density approximation (LDA) calculations \cite{oguchi,7} give a correct Fermi-surface topography, probed by de Haas-van Alphen measurements \cite{8}, as well as the magnetic enhancement due to Stoner exchange enhancement, although the mass renormalization cannot be explained within LDA calculations. In the superconducting state, the $^{101}$Ru nuclear spin lattice relaxation rate, $1/T_1$ exhibits a sharp decrease without a coherence peak (Hebel-Slichter peak) just above $T_c$, supporting the idea that an anisotropic pairing is effectively realized in Sr$_2$RuO$_4$ \cite{10}. In addition, the spontaneous appearance of an internal magnetic field below the transition temperature, reported by muon spin rotation measurements ($\mu SR$) \cite{11}, and the absence of $^{17}$O Knight shift modifications below $T_c$ \cite{ishida} point towards the triplet p-wave superconductivity. However, few experiments have really probed the exact nature of the spin fluctuations. Only the observation of a similar temperature dependence for $^{101}$Ru ${1/T_1T}$ and for $^{17}$O ${1/T_1T}$ in the NMR experiments by Imai {\it et al.} \cite{12} has suggested that spin fluctuations are predominantly FM in origin. The determination of the antiferromagnetic order in the closely related compound Ca$_2$RuO$_4$ \cite{caruo,nakatsuji} has suggested that the picture of a near-by FM instability in Sr$_2$RuO$_4$ is too simple. Furthermore, recent calculations which take into account the particular topology of the Fermi-surface, have predicted a sizeable magnetic response at the incommensurate wave-vector (2$\pi$/3a,2$\pi$/3a,0) \cite{9}, i.e. far away from the zone-center. The enhanced susceptibility arises from pronounced nesting properties of the almost one-dimensional d$_{xz,yz}$ bands. Mazin and Singh discuss the possibility of a competition between p-wave and d-wave superconductivity in Sr$_2$RuO$_4$ \cite{9}. In this letter, we report first inelastic neutron scattering (INS) measurements performed on single crystals of Sr$_2$RuO$_4$ in the normal state. Our data reveal dominant magnetic scattering at the incommensurate wave vectors {\bf$q_0$}=($\pm$0.6$\pi$/a,$\pm$0.6$\pi$/a,0), i.e. very close to the positions predicted by the band-structure calculations. The relevance of these findings for the mechanism of superconductivity in Sr$_2$RuO$_4$ will be discussed. Most of the INS measurements presented here have been carried out on a single crystal of cylindrical shape (4mm in diameter and 35mm long) grown by a floating zone method. The sample exhibits the superconducting transition at $T_c \sim$0.62 K. The single crystal was mounted in an aluminum can and attached to the cold finger of a closed cycle helium refrigerator. The INS experiments were performed on the triple axis spectrometers 2T (thermal beam) and 4F2 (cold beam) at the Laboratoire L\'eon Brillouin, Saclay, France. These spectrometers use neutron optics that focus the beam to the sample, with a resulting gain of neutron flux that proved to be crucial for these experiments. The experimental set up incorporates PG002 monochromator and analyzer and 14.7 meV fixed final energy. A pyrolytic graphite filter was inserted into the scattered beam in order to remove higher order contaminations. Data were taken within the scattering plane spanned by (1,0,0) and (0,1,0) directions. Some additional measurements were performed using several smaller single crystals with higher transition temperatures, $T_c$=1.4--1.5\ K; these experiments have revealed similar signals. Throughout this article, the wave vector {\bf Q}=(H,K,L) is indexed in units of the reciprocal tetragonal lattice vectors $2\pi /a=2\pi /b=1.63$ \AA$^{-1}$ and $2\pi /c=0.49$ \AA$^{-1}$ (I4/mmm space group)\cite{1}. \begin{figure}[t] \epsfxsize=6cm $$ \epsfbox{fig1.eps} $$ \caption{Constant-$\omega$ scans performed at $\hbar \omega$=6.2 meV around {\bf Q}=(1.3,0.3,0) along the (0,1,0) direction: T=10.4 K ($\bullet$), T=295 K ($\circ$).} \label{fig1} \end{figure} Figure \ref{fig1} shows representative constant-$\omega$ scans taken in the (H,K,0)-plane: at $\hbar \omega$=6.2 meV and around ${\bf Q_0} $ =(1.3,0.3,0) along the (0,1,0) direction. The scan at 10.4 K shows a sharp maximum of intensity peaked at ${\bf Q_0}$=(1.3,0.3,0) on top of a smooth background. At room temperature, this sharp peak has almost disappeared. The horizontal bar indicates the spectrometer resolution. At 10.4 K, several constant-$\omega$ scans, with 6.2 meV energy transfer and performed along different directions ((1,0,0), (0,1,0), (1,1,0), (1,-1,0)) have revealed the existence of comparable peaks at ${\bf Q}_0$ =${\bf q}_0$+${\bf G}$, where ${\bf q}_0$=($\pm$0.3,$\pm$0.3,0) $\equiv$ ($\pm$0.6$\pi$/a,$\pm$0.6$\pi$/a,0) and ${\bf G}$ is a zone-center or a Z-point (001) in the (HK0)-plane. The fit of the data to a Gaussian profile incorporating experimental resolution function demonstrates that the peak intensity is isotropic with an intrinsic q-width (FWHM), $\Delta q$=0.13 $\pm$ 0.02 \AA$^{-1}$. The interpretation of the scattering at ${\bf q}_0$ as magnetic in origin is supported by the large number of points in reciprocal space where it has been observed. Further, the lowest phonon frequencies at ${\bf q}_0$ are above 12 meV \cite{marcus2}. In addition, in contrast to a phonon-related scattering that increases at large $|Q|$ or with temperature, the scattering at ${\bf q}_0$ decreases both at large wave vector (Fig.~\ref{fig2}) and at high temperature (Fig.~\ref{fig1}). These different points establish the magnetic origin of the scattering observed around ${\bf q}_0$. In contrast, in spite of several attempts, no sizable FM spin fluctuations have been observed. \begin{figure}[tp] \epsfxsize=6cm $$ \epsfbox{fig2.eps} $$ \caption{ Magnetic intensity, measured at T=10.4 K and $\hbar \omega$=6.2 meV as a function of $|Q|$. For each point, the corresponding wave vector, (H,K,L), is also reported. The full line corresponds to the square of the $Ru^+$ magnetic form factor.} \label{fig2} \end{figure} In a paramagnetic state, the magnetic neutron cross section per formula unit can be written in terms of the imaginary part of the dynamical spin susceptibility, $\chi"({\bf Q}, \omega)$, as \cite{Lovesey,xirmq}; \begin{equation} \frac{d^2 \sigma}{d \Omega d \omega}=r_0^2 \frac{2 F^2({\bf Q})}{\pi (g \mu_B)^2} \frac{\chi"({\bf Q},\omega)}{1-\exp(-\hbar \omega / k_BT)} \; \label{eq:INS} \end{equation} where $r_0^2$=0.292 barn, $F({\bf Q})$ is the magnetic form factor and $g \simeq$2 is the Land\'e factor. The intensity of the scattering can be reasonably well described by the squared magnetic form factor of the Ru$^+$-ion \cite{formfac} (note that the magnetic form-factor of Ru$^{4+}$ is not available) after correction for geometrical factors related to the unfavorable shape of the sample, see Fig. 2. According to our measurements, the q-dependence of $\chi"$ is given by: $\chi"({\bf Q},\omega)= \chi"(q_0,\omega) \exp[-4\ln(2)( {\bf Q - Q_0})^2/\Delta q^2]$. \begin{figure}[tp] \epsfxsize=7cm $$ \epsfbox{fig3.eps} $$ \caption{Energy dependence of the imaginary part of the dynamical magnetic susceptibility at ${\bf Q}_0$=(1.3,0.3,0) as obtained from energy scans ($\circ$) and constant energy scans around ${\bf Q}_0$ along the (0,1,0) direction ($\bullet$) (see text).} \label{fig3} \end{figure} The Fermi surface in Sr$_2$RuO$_4$ is formed by three sheets \cite{9}: one, related to the $4d_{xy}$-orbitals is quasi-2D, whereas, the two others, related to $4d_{xz,yz}$ orbitals are quasi-1D. The 1D-sheets can be schematically described by parallel planes separated by $\bar q$=$\pm 2\pi /3a$, running both in the x and in the y directions. These peculiarities give rise to dynamical nesting effects at the wave vectors {\bf $k$}=($\bar q,k_y$), {\bf $k$}=($k_x,\bar q$) and in particular at ${\bf \bar q}$=($\bar q,\bar q$). The nesting effects become dominant when calculating the bare spin susceptibility of a non interacting metal \cite{9}, given by the Lindhard-function \cite{Lovesey}: \begin{equation} \chi_0 (q, \omega) = - 2 \mu_B^2 \sum_{k}\frac{f_{k+q}-f_{k}}{\varepsilon_{k+q}-\varepsilon_{k}-\hbar \omega + i \epsilon} \; \label{eq:3} \end{equation} where $\epsilon \rightarrow$0, $f_k$ is the Fermi distribution function and $\varepsilon_k$ the quasiparticle dispersion relation. Our INS are in very good agreement with the predicted four spots of magnetic scattering situated at ${\bf \bar q}$=($\pm$2$\pi$/3a,$\pm$2$\pi$/3a) \cite{9}. In the experiment the incommensurate magnetic responses are actually observed slightly away, at ${\bf q}_{0 //}$=($\pm0$.6$\pi$/a,$\pm0$.6$\pi$/a), which is most likely related to details of the band-structure \cite{9}. Let us now consider the energy dependence and magnitude of $\chi"(q_0,\omega)$. At T=10.4 K, constant-$\omega$ scans have been measured at {\bf Q}=(1.3,0.3,0) along the (0,1,0)-direction for different transferred energies between 2.4 and 12 meV. The magnetic response always displays a Gaussian profile, located at $q_0$ with an energy independent q-width, on top of a constant background. In addition, two energy scans have been performed at ${\bf Q}$=(1.3,0.3,0) and at {\bf Q}=(1.3,0.46,0), the latter providing a background reference. These measurements allow us to determine the energy dependence of the magnetic response at {\bf q$_0$} from 1.5 to 12 meV. The analysis could not be extended to higher and lower energies due to the contaminations by phonon\cite{marcus2} and elastic incoherent scattering respectively. Using Eq.~(\ref{eq:INS}), the magnetic intensity has been converted to the dynamical spin susceptibility $\chi$" after correction by the thermal population factor and the squared magnetic form factor reported in figure \ref{fig2}. We have then calibrated $\chi$" in absolute units against acoustic phonons, according to a standard procedure \cite{phonon}. $\chi "(\omega, Q_0)$, whose energy dependence is reported in absolute units in Fig.\ref{fig3}, slightly increases up to 7 meV and then almost saturates. This energy dependence can be parameterized following linear response theory: \begin{equation} \chi "(q_0, \omega) = \chi '(q_0, 0) \frac{\Gamma \omega}{\omega ^2 + \Gamma ^2} \; \label{eq:1} \end{equation} where $\Gamma$ is a damping energy of 9 meV and $\chi '(q_0, 0)$ = 180 $\mu_B^2.$eV$^{-1}$ corresponds to the static spin susceptibility at {\bf q$_0$}. It is worth emphasizing that $\chi '(q_0, 0)$ is 6 times larger than that at Q=0, i.e. the uniform susceptibility: $\tilde{\chi}=\chi '(Q=0, 0)$ = 30 $\mu_B^2.$eV$^{-1}$ ($\simeq$ 10$^{-3}$ emu/mole) \cite{1,2,3}. In La$_{1.86}$Sr$_{0.14}$O$_4$, usually referred to as a strongly correlated system, $\chi"({\bf Q}, \omega)$ at incommensurate wave vectors exhibits almost the same magnitude and a similar $\omega$-dependence \cite{gabe}. \begin{figure}[t] \vspace*{-0cm} \epsfxsize=6.3cm $$ \epsfbox{fig4.eps} $$ \caption[toto]{ Results from fits to a Gaussian profile of 6.2 meV constant-$\omega$ scans at ${\bf Q}_0$=(1.3,0.3,0) along the (0,1,0): temperature dependences of (a) $\chi"({\bf Q}_0,6.2 {\rm meV})$ and (b) the intrinsic q-width of the magnetic signal, $\Delta q$ (FWHM). (c) Comparison between $^{17}(1/T_{1}T)$ observed by $^{17}$O NMR by Imai {\it et al} \cite{12} ($\Box$) and the incommensurate contribution calculated from our INS measurements ($\bullet$). Assuming $\Lambda$=33 kOe/$\mu_B$\cite{Aq2} the two scales in this figure are identical. } \label{fig4} \end{figure} In Sr$_2$RuO$_4$, electronic correlations are incorporated in RPA calculations: the spin susceptibility $\chi (q, \omega)$ becomes enhanced through the Stoner-factor $I(q)$\cite{7,9}: \begin{equation} \chi (q, \omega) = \frac{\chi_0(q, \omega)}{1 - \frac{I(q)}{2 (\mu_B)^2} \chi_0(q, \omega)} \; \label{eq:2} \end{equation} The q-dependence of the Stoner factor, for an individual RuO$_2$ plane, reflects the fact that FM interactions are favored over antiferromagnetic interactions in Sr$_2$RuO$_4$ : in our units, $I(q)= 0.86/(1+0.8(a/\pi)^2q^2)$ eV (q in \AA$^{-1}$)\cite{7,9}. INS results point towards a strong enhancement of the spin susceptibility by the Stoner factor (see Eq.~(\ref{eq:2})), such that the system should be close to a magnetic instability at ${\bf q}_0$. With $\chi '(q_0, 0)$ = 180 $\mu_B^2.$eV$^{-1}$, one deduces from Eq.~(\ref{eq:2}) that $\frac{I(q_0)}{2(\mu_B)^2}\chi_0(q_0, 0) \simeq$0.99, instead of being larger than 1 for a magnetic instability. Thus, incommensurate spin fluctuations are stronger than FM fluctuations in Sr$_2$RuO$_4$, as suggested in ref.~\cite{9}. The temperature dependence of both $\chi "({\bf Q}_0, 6.2 {\rm meV})$ and the intrinsic q-width are reported in Fig.~\ref{fig4}, as deduced from constant-$\omega$ scans performed at 6.2 meV around {\bf Q}=(1.3,0.3,0) along the (0,1,0) direction at different temperatures. $\chi "({\bf q}_0, 6.2 {\rm meV})$ exhibits a sharp decrease upon temperature increase and simultaneously the magnetic response broadens (the width of the signal can be reliably determined only up to 200 K). The T-dependence of $\chi"$ observed in INS measurements may be described by the out-smearing of the Fermi surface due to thermal hopping of electrons into unoccupied states (see the numerator in Eq.~(\ref{eq:3})), yielding a lowering of the dynamical susceptibility at ${\bf q}_0$ and its broadening in q-space. The T-dependence of the magnetic response at ${\bf q}_0$ can indeed be qualitatively reproduced \cite{pfeuty} using Eqs. (\ref{eq:3})-(\ref{eq:2}) and a description of the LDA band structure by three mutually non-hybridizing tight-binding bands \cite{7}. INS measurements point out the existence of strong magnetic response at ${\bf q}_0$, but do not reveal any sizeable FM fluctuations. In contrast, the uniform spin susceptibility \cite{1,2,3} and the Knight shift measurements \cite{12,13} provide evidence of strong FM correlation in Sr$_2$RuO$_4$. However, the delicate balance between FM and incommensurate spin fluctuations should become visible in the spin-lattice relaxation rate $T_1$ measured by both $^{17}$O and $^{101}$Ru NMR experiments\cite{12,13}. These NMR-techniques probe the low energy spin fluctuations ($\omega \rightarrow 0$ with respect to INS measurements); furthermore, they integrate the fluctuations in q-space. Since the INS studies have determined the incommensurate fluctuations on an absolute scale we may estimate their contribution to $(1/T_1T)$, $^{INS} (1/T_1T)$. In general $(1/T_1T)$ probes the q-summation of the the imaginary part of the susceptibility divided by the frequency in the limit $\omega \rightarrow 0$ (i.e. its initial slope), ${\sum_{q} \frac{\chi "(q,\omega)}{ \omega} }\vert_{\omega \rightarrow 0}$; its temperature dependency is shown in Fig. 4.c (left scale). What renders the quantitative comparison between the INS and NMR-results more difficult is the estimate of the hyperfine field whose q-dependent Fourier transform, $A(q)$, weights the susceptibility in NMR-studies. Considering that INS magnetic fluctuations are sharply peaked around {\bf q$_0$}, one may approximate $A(q)$=$A(q_0)$, and gets\cite{berthier}, \begin{equation} ^{ INS}(1/T_1T) \simeq \frac{k_B\gamma_n^2}{(g\mu_B)^2} \left. | A(q_0) |^2 {\sum_{q} \frac{\chi "(q,\omega)}{ \omega} }\right\vert_{\omega \rightarrow 0} \; \label{eq:1/T1} \end{equation} with $| A(q_0) |^2= \Lambda ^2 [1+1/2(\cos(2\pi 0.3)+\cos(2\pi / a0.3))]$ ($\Lambda$ = 33 kOe/${\mu _B}$\cite{Aq2}) for $^{17}$O and $A(q_0)$=-299 kOe/$\mu_B$\cite{10,12} for $^{101}$Ru. Using these values, we directly compare $^{INS}(1/T_1T)$ with the measured $^{17}$O $(1/T_1T)$ in Fig. \ref{fig4}.c (right scale). Clearly, the spin fluctuations at ${\bf q}_0$ significantly contribute to $^{17}(1/T_1T)$, and can explain a large part of the reported T-dependence\cite{12,13}. Similar calculation for the $^{101}$Ru $(1/T_1T)$ (not shown) yields even a stronger contribution. The remaining parts in $^{17,101}(1/T_1T)$, likely associated with FM excitations, should exhibit a less pronounced T-dependence similar to that of the uniform static spin susceptibility. Furthermore, assuming a weak q-dependence for these FM excitations\cite{pfeuty}, the comparison of NMR and INS measurements allows us to estimate the ferromagnetic characteristic energy to be of the order of 50 meV. This rather elevated value actually provides a satisfactory explanation for the absence of FM fluctuations in our INS measurements. To conclude, our INS measurements demonstrate the existence of incommensurate spin fluctuations related to dynamical nesting properties of the Sr$_2$RuO$_4$ Fermi surface. Our data suggest that the system is close to a magnetic instability at ${\bf q}_{0//}$=($\pm 0.6 \pi /a$,$\pm 0.6 \pi /a$). The comparison of INS and $^{17}(1/T_1T)$ measurements suggests that the FM fluctuations are transferred to higher energy with respect to the spin fluctuations at ${\bf q}_0$. All these results cast some doubt on the predominant role of FM spin fluctuations in the mechanism of superconductivity in Sr$_2$RuO$_4$. We wish to acknowledge P. Pfeuty, J. Bobroff and Ph. Mendels for helpful discussions.
\section{Introduction} \noindent B\"{a}cklund transformations (BTs) are an important aspect of the theory of integrable systems which have traditionally been studied in the context of evolution equations. However, more recently there has been much interest in discrete systems or integrable mappings \cite{sympl, veselov}. Within the modern approach to separation of variables (reviewed by Sklyanin in \cite{sklyan}) this has led to the study of BTs for finite-dimensional Hamiltonian systems \cite{kuskly}. The latter are canonical transformations including a B\"{a}cklund parameter $\lambda$, and apart from being interesting integrable mappings in their own right they also lead to separation of variables when $n$ such mappings are applied to an integrable system with $n$ degrees of freedom. The sequence of B\"{a}cklund parameters $\lambda_{j}$ together with a set of conjugate variables $\mu_{j}$ constitute the separation variables, and satisfy a new property called {\it spectrality} introduced in \cite{kuskly}. We proceed to develop these ideas with some new examples of BTs for $n$-body systems, namely the many-body generalisation of the case (ii) integrable H\'{e}non-Heiles system, the Garnier system and the Neumann system on the sphere (see \cite{eekt}). It is known that the case (ii) H\'{e}non-Heiles system is equivalent to the stationary flow of fifth-order KdV \cite{fordy}, while the Garnier and Neumann systems may be obtained as restricted flows of the KdV hierarchy \cite{zeng}. Thus we derive BTs for these systems by reduction of the standard BT for KdV, which arises from the Darboux-Crum transformation \cite{crum} for Schr\"{o}dinger operators. The restriction of the Darboux transformation to the stationary flows of the modified (mKdV) hierarchy has been discussed in \cite{poiss}. In the following section we describe how the reduction works in general, before specialising these considerations to each particular system and presenting the associated generating function for the BT. We note that these systems are examples of the reduced Gaudin magnet \cite{eekt}, so that we have the following Lax matrix \begin{equation} L(u)=\sum_{j=1}^{n}\frac{\ell_{j}}{u-a_{j}}+B(u), \qquad \ell_{j}=\left( \begin{array}{cc} S_{j}^{3} & S_{j}^{-} \\ S_{j}^{+} & -S_{j}^{3} \end{array} \right) \label{eq:gaulax} \end{equation} where (up to scaling) the $S_{j}$ satisfy $n$ independent copies of the standard $sl(2)$ algebra: \begin{equation} \{ S_{j}^{3},S_{k}^{\pm} \} =\pm2\delta_{jk}S_{k}^{\pm}, \qquad \{ S_{j}^{+},S_{k}^{-} \} =4\delta_{jk}S_{k}^{3}. \label{eq:alg} \end{equation} For the H\'{e}non-Heiles and Garnier systems the matrix $B(u)$ is respectively quadratic and linear in the spectral parameter $u$, while for the Neumann system it is independent of $u$ and turns out to be constant due to the constraint that the particles lie on the sphere (hence the Poisson algebra (\ref{eq:alg}) must be modified by Dirac reduction). We have constructed the BT for the (non-reduced) $sl(2)$ Gaudin magnet with quasi-periodic boundary condition in \cite{gaud}, while some preliminary results on the classical Garnier system and two-body H\'{e}non-Heiles system first appeared in \cite{us}. \section{Classical integrable systems and KdV} \setcounter{equation}{0} \subsection{Restricting the BT} \noindent As is well known, the Darboux-Crum transformation \cite{crum} consists of mapping the Schr\"{o}dinger operator $\partial_{t}^{2}+V-\lambda$ to another operator $\partial_{t}^{2}+\tilde{V}-\lambda$ by factorizing the former and then reversing the order of factorisation. Given an eigenfunction $\phi$ satisfying $$ (\partial_{t}^{2}+V-\lambda)\phi=0 $$ we may set $y=(\log[\phi])_{t}$ and then \begin{equation} V=-y_{t}-y^{2}+\lambda, \qquad \tilde{V}=y_{t}-y^{2} +\lambda; \label{eq:miura} \end{equation} for $\lambda=0$ this is just the Miura map for KdV. Also given another eigenfunction $\psi$ of the Schr\"{o}dinger operator with potential $V$ for a different spectral parameter $u$ we have $$ (\partial_{t}^{2}+V-u)\psi=0, \qquad (\partial_{t}^{2}+\tilde{V}-u)\tilde{\psi}=0 $$ where the transformation to the new eigenfunction $\tilde{\psi}$ and its derivative may be given in matrix form as \begin{equation} \left(\begin{array}{c} \tilde{\psi}_{t} \\ \tilde{\psi} \end{array} \right) = k \left(\begin{array}{cc} -y & y^{2} + u - \lambda \\ 1 & -y \end{array} \right) \left(\begin{array}{c} \psi_{t} \\ \psi \end{array} \right) \label{eq:gauge} \end{equation} for any constant $k$. From (\ref{eq:miura}) follows the standard formula for the Darboux-B\"{a}cklund transformation of KdV, $\tilde{V}=V+2(\log[\phi])_{tt}$. For what follows it will also be necessary to consider a product of eigenfunctions for a Schr\"{o}dinger operator with potential $V$ and eigenvalue $u$, $$ f=\psi\psi' $$ with Wronskian $\psi_{t}\psi '-\psi\psi_{t}'=2m$. It is well known that $f$ satisfies the Ermakov-Pinney equation \cite{pinney} \begin{equation} ff_{tt}-\frac{1}{2}f_{t}^{2}+2(V-u)f^{2}+2m^{2}=0. \label{eq:pinney} \end{equation} If we now transform $\psi$ and $\psi'$ according to (\ref{eq:gauge}) then we find a new product of eigenfunctions $\tilde{f}=\tilde{\psi} \tilde{\psi}'$ satisfying the same Ermakov-Pinney equation but with $V$ replaced by $\tilde{V}$, given explicitly by \begin{equation} \tilde{f}=(\lambda-u)^{-1}\frac{(Z^{2}-m^{2})}{f}, \quad Z=\frac{1}{2}f_{t}-yf, \label{eq:trans} \end{equation} where we have set $k^{2}=(\lambda-u)^{-1}$ to ensure that the transformed eigenfunctions have the same Wronskian $2m$. It is also straightforward to show that, in terms of $\tilde{f}$, the quantity $Z$ can be written as $Z=-\frac{1}{2}\tilde{f}_{t}-y\tilde{f}$ (see \cite{exact}). We can now describe how this transformation restricts to the finite-dimensional Hamiltonian systems presented below. The systems are expressed in variables $(q_{j},p_{j})$ which appear in the Lax matrix (\ref{eq:gaulax}) via the identification \cite{ku,eekt} $$ S_{j}^{3}=p_{j}q_{j}, \quad S_{j}^{-}=-p_{j}^{2}+\frac{m_{j}^{2}}{q_{j}^{2}}, \quad S_{j}^{+}=q_{j}^{2}. $$ For H\'{e}non-Heiles and Garnier the non-vanishing Poisson brackets are the standard ones $\{p_{j},q_{k}\}=\delta_{jk}$ which provide a realization of the algebra (\ref{eq:alg}); for the Neumann system on the sphere the bracket must be modified by Dirac reduction. All of the systems are Liouville integrable, and thus have a complete set of Hamiltonians in involution, but for these purposes we concentrate on the Hamiltonian $h$ generating the flow corresponding to $t$ above (in KdV theory this is usually denoted $x$, the spatial variable). For this flow the Lax equation $L_{t}=[N,L]$ is the compatibility condition for the linear system \begin{equation} L(u)\Psi=v\Psi, \quad \Psi_{t}=N\Psi; \quad N=\left(\begin{array}{cc} 0 & u-V(q_{j},p_{j}) \\ 1 & 0 \end{array} \right). \label{eq:linsys} \end{equation} Observe that the second part of the linear system is just a Schr\"{o}dinger equation for the potential $V$; for Neumann and Garnier this is a function of $(q_{j},p_{j})$ for $j=1,\ldots,n$, while for H\'{e}non-Heiles there is an extra pair of conjugate variables $(q_{n+1},p_{n+1})$ such that $V\equiv q_{n+1}$. The equations of motion generated by this Hamiltonian take the form $q_{j,t}=p_{j}$ and \begin{equation} p_{j,t}=q_{j,tt}=(a_{j}-V(q_{k},p_{k}))q_{j}- \frac{m_{j}^{2}}{q_{j}^{3}} \label{eq:pini} \end{equation} for $j=1,\ldots,n$; for H\'{e}non-Heiles there are also equations for $q_{n+1}$ and $p_{n+1}=q_{n+1,t}$. The important thing to observe is that (\ref{eq:pini}) is equivalent to the fact that $S_{j}^{+}=q_{j}^{2}$ satisfies the Ermakov-Pinney equation (\ref{eq:pinney}) corresponding to a Schr\"{o}\-dinger equation with potential $V$ and eigenvalue $a_{j}$. Thus to obtain a B\"{a}cklund transformation for these many-body systems we simply apply a Darboux-Crum transformation (\ref{eq:miura}) to the potential $V=V(q_{j},p_{j})$ to obtain $\tilde{V}=V(\tilde{q}_{j},\tilde{p}_{j})$, and then we know that the solutions of the Ermakov-Pinney equation must transform according to (\ref{eq:trans}). By this procedure we may explicitly construct the BT for the many-body systems below (or for any restricted flow of KdV), and it is then simple to calculate the generating function $F(q_{j},\tilde{q}_{j})$ of this canonical transformation, such that $$ dF=\sum_{j} (p_{j}dq_{j}-\tilde{p}_{j}d\tilde{q}_{j}). $$ The discrete Lax equation for the BT, $$ ML=\tilde{L}M $$ where $\tilde{L}=L(\tilde{q_{j}}, \tilde{p_{j}};u)$, is necessary to ensure the preservation of the spectral curve $\det(v-L(u))=0$ (so that all the Hamiltonians in involution are preserved). This follows immediately from the properties of the Darboux-Crum transformation, since we know that the vector $\Psi$ in the linear system (\ref{eq:linsys}) must transform according to (\ref{eq:gauge}), and hence we may take (setting $k=1$) \begin{equation} M=\left( \begin{array}{cc} -y & y^{2} + u - \lambda \\ 1 & -y \end{array} \right). \label{eq:emm} \end{equation} Of course we must determine $y$ as a function of the dynamical variables. In the Garnier and H\'{e}non-Heiles cases it turns out that the potential depends on coordinates only, $V=V(q_{j})$, and so by adding the two equations in (\ref{eq:miura}) we obtain $$ y(q_{j}, \tilde{q_{j}})=\pm \sqrt{\lambda-\frac{1}{2}(V+\tilde{V})}; $$ to obtain the correct continuum limit of the discrete dynamics it is necessary to take the negative branch of the square root (see \cite{us, exact}). For the Neumann system $V$ depends on both coordinates and momenta, so the above does not yield $y(q_{j}, \tilde{q_{j}})$. There is another way of writing $L$ which arises more naturally via reduction from the zero curvature representation of the KdV hierarchy \cite{eekt,fordy, zeng}, viz $$ L(u)=\left(\begin{array}{cr} \frac{1}{2}\Pi_{t} & -\frac{1}{2}\Pi_{tt} +(u-V)\Pi\\ \Pi & -\frac{1}{2}\Pi_{t} \end{array} \right) $$ where \begin{equation} \Pi(u)=\sum_{j=1}^{n}\frac{q_{j}^{2}}{u-a_{j}} +\Delta(u). \label{eq:other} \end{equation} $\Delta$ is a polynomial in $u$ fixing the dynamical term $B$ in (\ref{eq:gaulax}); we shall present the appropriate $\Delta$ and $B$ in each case below. Clearly the $t$ derivatives of $\Pi$ can be rewritten using the equations of motion to yield (\ref{eq:gaulax}). Finally if we write the (hyper-elliptic) spectral curve as $$ v^{2}=R(u) $$ then it is easy to check that the spectrality property \cite{kuskly} is satisfied for these systems, in the sense that defining the conjugate variable to $\lambda$ by $$ \mu=-2\frac{\partial F}{\partial \lambda} $$ we find that $$ L(\lambda)\Omega=\mu\Omega $$ with eigenvector $\Omega=(y,1)^{T}$, or in other words $\mu^{2}=R(\lambda)$ so that $(\lambda,\mu)$ is a point on the spectral curve. Note that (as for the examples in \cite{gaud, kuskly}) this eigenvector spans the kernel of $M$, $$ M(\lambda)\Omega=0. $$ We can also write $y$ explicitly in terms of both the old and the new variables related by the BT, thus: \begin{equation} y(q_{j},p_{j})= \frac{\Pi_{t}(\lambda)+2\mu}{2\Pi(\lambda)}, \qquad y(\tilde{q_{j}},\tilde{p_{j}}) = -\frac{(\tilde{\Pi}_{t}(\lambda)-2\mu)} {2 \tilde{\Pi}(\lambda)}; \label{eq:ynewold} \end{equation} clearly we denote $\tilde{\Pi}(\lambda)= \Pi(\tilde{q_{j}},\tilde{p_{j}};\lambda)$. \subsection{Generalised H\'{e}non-Heiles system} \noindent For the many-body generalisation of case (ii) integrable H\'{e}non-Heiles system, the Hamiltonian generating the $t$ flow takes the form $$ h=\frac{1}{2}\sum_{j=1}^{n+1}p_{j}^{2} +q_{n+1}^{3}+q_{n+1} \left(\frac{1}{2}\sum_{j=1}^{n} q_{j}^{2}+c\right) -\frac{1}{2}\sum_{j=1}^{n} \left(a_{j}q_{j}^{2}+ \frac{m_{j}^{2}}{q_{j}^{2}}\right). $$ The original case (ii) integrable H\'{e}non-Heiles system corresponds to $n=1$ with $c=m_{j}=a_{j}=0$. The link between stationary fifth-order KdV and the type (ii) system was noted by Fordy in \cite{fordy}, although this was anticipated in work of Weiss \cite{weiss}, who used Painlev\'{e} analysis to derive a BT and associated linear problem (a similar result also appears in \cite{new}). None of these authors wrote a BT explicitly as a canonical transformation with parameter, although (without parameter) this was done for a non-autonomous version in \cite{thesis}. For the Lax matrix $L$ of the generalised $(n+1)$-body H\'{e}non-Heiles system we have $\Delta=-16u-8q_{n+1}$ so that the extra term $B(u)$ is given by $$ B=\left(\begin{array}{cc} -4p_{n+1} & E \\ -16u-8q_{n+1} & 4p_{n+1} \end{array}\right), \quad E=-16u^{2}+8q_{n+1}u -4q_{n+1}^{2}-\sum_{j=1}^{n} q_{j}^{2}-4c. $$ The equations of motion for $h$ imply that the squares of the first $n$ coordinates $q_{j}^{2}$ satisfy the Ermakov-Pinney equation (\ref{eq:pinney}) for $m=m_{j}$ with $$ V=q_{n+1} $$ and eigenvalue $a_{j}$. Thus the BT for the system can be calculated directly by applying the Darboux-Crum transformation to $V=q_{n+1}$, to yield $\tilde{V}=\tilde{q}_{n+1}$, and applying (\ref{eq:trans}) to each $q_{j}^{2}$ for $j=1,\ldots,n$. After some calculation the generating function for this canonical transformation is found to be $$ F(q_{j},\tilde{q}_{j};\lambda)=\sum_{j=1}^{n} \left( Z_{j}+\frac{m_{j}}{2}\log\left[ \frac{Z_{j}-m_{j}}{Z_{j}+m_{j}}\right]\right) + \frac{16}{5}y^{5}+4(q_{n+1}+\tilde{q}_{n+1})y^{3} $$ $$ +\left(2q_{n+1}^{2}+2q_{n+1}\tilde{q}_{n+1} +2\tilde{q}_{n+1}^{2}+\frac{1}{2}\sum_{j=1}^{n} (q_{j}^{2}+\tilde{q}_{j}^{2})+2c\right)y, $$ where we have found it convenient to use the quantities $Z_{j}(q_{j},\tilde{q}_{j})$ and $y(q_{j},\tilde{q}_{j})$ defined by \begin{equation} Z_{j}^{2}=m_{j}^{2}+ (\lambda-a_{j})q_{j}^{2}\tilde{q}_{j}^{2}, \label{eq:zeb} \end{equation} and $$ y=-\sqrt{\lambda-\frac{1}{2}(q_{n+1}+\tilde{q}_{n+1})}. $$ In order to check the spectrality property, we have explicitly found that the eigenvalue of $L(\lambda)$ with eigenvector $\Omega=(y,1)^{T}$ can be written as $$ \mu(q_{j},\tilde{q}_{j};\lambda) =-\sum_{j=1}^{n}\frac{Z_{j}}{\lambda-a_{j}}- \frac{1}{y}\frac{\partial F}{\partial y}, $$ which precisely equals $-2\frac{\partial F}{\partial \lambda}$ as required. \subsection{Garnier system} \noindent For the Garnier system the $t$ flow is generated by the Hamiltonian $$ h=\frac{1}{2}\sum_{j=1}^{n}p_{j}^{2} +\frac{1}{2}(\sum_{j=1}^{n}q_{j}^{2})^{2} -\frac{1}{2}\sum_{j=1}^{n}\left(a_{j}q_{j}^{2}+ \frac{m_{j}^{2}}{q_{j}^{2}}\right). $$ This differs from the traditional Garnier system as in \cite{us, wojc, zeng} by the inclusion of extra inverse square terms. The Newton equations for the $q_{j}$ are $$ q_{j,tt}+2(\sum_{k}q_{k}^{2})q_{j}= a_{j}q_{j}-\frac{m_{j}^{2}}{q_{j}^{3}}, $$ so clearly for the standard restricted flows of KdV \cite{zeng}, when $m_{j}=0$, each $q_{j}$ is an eigenfunction of a Schr\"{o}dinger operator with potential $$ V=2\sum_{j}q_{j}^{2} $$ and eigenvalue $a_{j}$, while in general $q_{j}^{2}$ is a product of eigenfunctions satisfying the Ermakov-Pinney equation \cite{pinney} for $m=m_{j}$. The Lax matrix of the Garnier system has $\Delta=1$, so $L$ takes the form (\ref{eq:gaulax}) with $$ B=\left(\begin{array}{cc} 0 & u-\sum_{j}q_{j}^{2} \\ 1 & 0 \end{array}\right). $$ Applying the Darboux-Crum transformation we obtain a new potential $$ \tilde{V}=2\sum_{j}\tilde{q}_{j}^{2}, $$ and the corresponding BT induced on the Garnier system is equivalent to gauging $L$ by the matrix $M$ of the form (\ref{eq:emm}) with $$ y=-\sqrt{\lambda-\sum_{j} (q_{j}^{2}+\tilde{q}_{j}^{2})}. $$ Finally we can calculate the generating function for this BT, which may be written as follows: $$ F(q_{j},\tilde{q}_{j};\lambda)=\sum_{j=1}^{n} \left( Z_{j}+\frac{m_{j}}{2}\log\left[ \frac{Z_{j}-m_{j}}{Z_{j}+m_{j}}\right]\right) -\frac{1}{3}y^{3}, $$ where $y(q_{j},\tilde{q}_{j})$ is as above and $Z_{j}$ is given by the same expression (\ref{eq:zeb}) as for H\'{e}non-Heiles. In \cite{us} we derived this generating function for the special case $m_{j}=0$ when the logarithm terms do not appear. To check spectrality we notice that $L(\lambda)$ has eigenvalue $$ \mu(q_{j},\tilde{q}_{j};\lambda) =-\sum_{j=1}^{n}\frac{Z_{j}}{\lambda-a_{j}} +y $$ with eigenvector $\Omega$, and so we see that $\mu=-2\frac{\partial F}{\partial \lambda}$. \subsection{Neumann system on the sphere} \noindent For the Neumann system the $t$ flow is generated by $$ h=\frac{1}{2}\sum_{j=1}^{n}p_{j}^{2} -\frac{1}{2}\sum_{j=1}^{n}\left(a_{j}q_{j}^{2}+ \frac{m_{j}^{2}}{q_{j}^{2}}\right). $$ Once again this has extra inverse square terms compared with the standard Neumann system \cite{ragn, wojc}. The Poisson bracket for this system is modified by constraining the particles to lie on a sphere, so that \begin{equation} (q,q)\equiv\sum_{j}q_{j}^{2}=const, \qquad (q,p)\equiv\sum_{j}q_{j}p_{j}=0 \label{eq:constr} \end{equation} which results in the non-vanishing Dirac brackets \begin{equation} \{p_{j},q_{k}\}=\delta_{jk}- \frac{q_{j}q_{k}}{(q,q)}, \quad \{p_{j},p_{k}\}=\frac{q_{j}p_{k}-q_{k}p_{j}}{(q,q)}. \label{eq:dirac} \end{equation} With this bracket the Hamilton equations are $q_{j,t}=p_{j}$ and (\ref{eq:pini}) with $$ V=(q,q)^{-1} \sum_{j}\left( p_{j}^{2}+a_{j}q_{j}^{2} -\frac{m_{j}^{2}}{q_{j}^{2}}\right). $$ The Lax matrix for the Neumann system arises by setting $\Delta=0$, which in (\ref{eq:gaulax}) gives the following matrix $B$: $$ B=\left(\begin{array}{cc} 0 & (q,q) \\ 0 & 0 \end{array}\right). $$ In fact if we start from the linear system (\ref{eq:linsys}) and leave $V$ unspecified then (\ref{eq:pini}) as well as the constraint $(q,q)_{t}=0$ are consequences of the Lax equation, and together these are sufficient to determine the form of $V$; this is also how the equations for the constrained Neumann system arise in a Lagrangian approach \cite{ragn}. Given that the phase space is now degenerate with two Casimirs given by (\ref{eq:constr}), it would appear that the standard sort of generating function will no longer be appropriate for describing a BT. It turns out that we can apply the Darboux-Crum transformation as before, and transform the quantities $q_{j}^{2}$ according to (\ref{eq:trans}). In this way we obtain new variables $\tilde{q}_{j}(q_{k},p_{k})$ and $\tilde{p}_{j}(q_{k},p_{k})$, which are naturally written with the use of the quantity $y(q_{k},p_{k})$ given by the first formula in (\ref{eq:ynewold}); on the Lax matrix this transformation arises by gauging with $M$ as in (\ref{eq:emm}). Similarly the transformation can be inverted to give $q_{j}(\tilde{q}_{k},\tilde{p}_{k})$ and $p_{j}(\tilde{q}_{k},\tilde{p}_{k})$ written in terms of $y(\tilde{q}_{k},\tilde{p}_{k})$ given by the right hand formula of (\ref{eq:ynewold}). However, it would still be nice to write a generating function for this transformation. We have found that if we formally take $$ F(q_{j},\tilde{q}_{j};\lambda)=\sum_{j=1}^{n} \left( Z_{j}+\frac{m_{j}}{2}\log\left[ \frac{Z_{j}-m_{j}}{Z_{j}+m_{j}}\right] +\frac{1}{2}y(q_{j}^{2}-\tilde{q}_{j}^{2})\right) $$ with $Z_{j}$ given by (\ref{eq:zeb}) as usual, and regard $y$ as a sort of Lagrange multiplier (independent of the coordinates and $\lambda$), then we do indeed obtain the correct expressions $$ p_{j}=\frac{\partial F}{\partial q_{j}}, \qquad \tilde{p}_{j}=-\frac{\partial F}{\partial\tilde{q}_{j}}, $$ but these contain $y$ which is unspecified. If we then require that the constraints (\ref{eq:constr}) are preserved under the BT applied from old to new variables or vice-versa, then in either direction the constraints are preserved if and only if $y$ satisfies a quadratic equation with solution given respectively by the formulae (\ref{eq:ynewold}). Alternatively if we require spectrality then second component of the equation $L(\lambda)\Omega=\mu\Omega$ gives $$ \mu(q_{j},\tilde{q}_{j};\lambda) =-\sum_{j=1}^{n}\frac{Z_{j}}{\lambda-a_{j}}= -2\frac{\partial F}{\partial \lambda} $$ as required, while the first component gives (after making use of the formula (\ref{eq:zeb}) and the BT) $$ \mu=-\sum_{j=1}^{n}\frac{Z_{j}}{\lambda-a_{j}} +\frac{1}{y}\sum_{j}(q_{j}^{2}-\tilde{q}_{j}^{2}). $$ Hence spectrality requires that the second term vanishes, and so the first constraint (\ref{eq:constr}) is preserved; the preservation of the second constraint is then an algebraic consequence of the BT. Thus we see that for this BT we can write the new variables as functions of the old and vice-versa, but a formula for $y(q_{j},\tilde{q}_{j};\lambda)$ is lacking. Also this discretization of the Neumann system is apparently new, since it is exact (preserving the Lax matrix for the continuous system) unlike the Veselov or Ragnisco discretizations discussed in \cite{ragn}. \section{Conclusions} \noindent It would also be interesting to look at BTs with parameter in the non-autonomous case \cite{thesis}, where deformation with respect to the B\"{a}cklund parameter would probably have to be introduced (corresponding to the associated isomonodromy problem). \section{Acknowledgements} \noindent ANWH thanks the Leverhulme Trust for providing a Study Abroad Studentship in Rome, and is grateful to J.~Harnad and Y.~Suris for useful conversations. VBK acknowledges the support from the EPSRC and the support from Istituto Nazionale di Fisica Nucleare for his visit to Rome. The authors would also like to thank the organisers of the meeting {\it Integrable Systems: Solutions and Transformations} in Guardamar, Spain (June 1998) where some of this work was carried out.
\section{Introduction} \medskip Supersymmetry provides a promising solution to the gauge hierarchy problem afflicting the standard model (SM). However, it is clear that supersymmetry must be broken at low energies. The specific mechanism for transmitting supersymmetry breaking effects is important in determining the low-energy experimental signatures. Currently, there are two known ways that supersymmetry breaking effects appear in the low-energy Lagrangian. In gravity-mediated scenarios~\cite{grav}, supersymmetry is broken in a hidden sector and transmitted gravitationally to the observable sector fields. While this scenario is elegant and simple, it suffers from the supersymmetric flavor problem. Alternatively, in gauge-mediated scenarios~\cite{gaugemed}, supersymmetry breaking is transmitted via gauge forces and this scenario provides an appealing solution to the supersymmetric flavor problem. Both of these alternative scenarios have distinct experimental signatures. We consider a third scenario for transmitting supersymmetry breaking to the observable sector. In this scenario, rescaling anomalies in the supergravity Lagrangian give rise to soft mass parameters for the observable sector fields~\cite{randall,giudice}. Unlike the gravity-mediated or gauge-mediated scenarios, these anomaly contributions will always be present if supersymmetry is broken. We will refer to the case in which the anomaly-induced masses are dominant as the anomaly-mediated supersymmetry breaking (AMSB) scenario. In this scenario the gaugino mass is proportional to the corresponding gauge beta function while the scalar masses (and $A$-terms) depend on the anomalous dimensions of the corresponding scalar fields. One of the distinctive features of the AMSB scenario is the gaugino mass spectrum, with the Wino being the lightest supersymmetric particle. Similarly, the squark mass spectrum is unique but unfortunately the slepton mass spectrum is tachyonic. This can be cured by adding a positive, non-anomaly mediated contribution~\cite{randall}. Some phenomenological consequences of this scenario have been recently presented in ref.~\cite{feng}. A different and very interesting approach to cure the tachyonic mass spectrum problem has been suggested in ref.~\cite{pr}. A distinctive feature of AMSB is that the gravitino is much heavier than the gauginos and squarks. This is cosmologically attractive because the gravitino problem can be ameliorated. Moreover, gravitino decays can produce a present Wino energy density close to the critical value. The neutral Wino is therefore a good dark-matter candidate, in spite of its negligible thermal relic density. \section{The anomaly-induced mass spectrum} The anomaly-induced soft terms~\cite{randall,giudice} are always present in a broken supergravity theory, regardless of the specific form of the couplings between the hidden and observable sectors. They are linked to the existence of the superconformal anomaly. Indeed they explicitly arise when one tries to eliminate from the relevant Lagrangian the supersymmetry-breaking auxiliary background field by making a suitable Weyl rescaling of the superfields in the observable sector. Their origin has been discussed from various point of views in refs.~\cite{randall,giudice,pr}. Here we give a heuristic derivation of the essential results, and make some comments on their phenomenological relevance. The effect of supersymmetry breaking can be described by a flat-space chiral superfield $\Phi$, with background value \begin{equation} \Phi =1-m_{3/2} \theta^2. \label{phiback} \end{equation} This field acts as a compensator of the super-Weyl transformation. In other words, by choosing suitable couplings of $\Phi$ to the observable fields, the theory is made superconformal invariant. Let us consider a supersymmetric gauge theory with no mass parameters at the classical level. This does not appear at first sight to be relevant to the minimal supersymmetric model which contains a mass term -- the Higgs mixing mass $\mu$ -- seemingly even in the limit of exact supersymmetry. Actually, the $\mu$ term can be viewed as an effect of supersymmetry breaking~\cite{gmas}, and therefore we set it to zero for the moment. Mechanisms for generating $\mu$ in AMSB scenarios have been discussed in refs.~~\cite{randall,giudice,pr}. At the quantum level, there is always the need to introduce a mass parameter, which is the renormalization scale $\mu$ (not to be confused with the Higgs mixing parameter). In the presence of a compensator field $\Phi$ for super-Weyl transformations, it is natural to expect that the renormalization scale $\mu$ is promoted to a superfield, according to \begin{equation} \mu \rightarrow \mu / \sqrt{\Phi^\dagger \Phi} . \label{cont} \end{equation} The replacement of $\Phi$ with its background value given in Eq.~(\ref{phiback}) generates a specific set of supersymmetry-breaking terms. The simplest way to obtain the form of the supersymmetry-breaking terms is to employ the technique developed in ref.~\cite{gr}. The main idea is that when certain parameters of a supersymmetric theory are ``analytically continued" into superspace, the renormalization-group (RG) flow of the modified theory is completely determined by the properties of the original theory. In particular, if a parameter is continued into a supersymmetry-breaking background field, the RG properties of the exact supersymmetric theory determine the form of the soft terms. The prescription given in Eq.~(\ref{cont}) is a specific example of such a continuation. We can then make use of the general expressions of the gaugino masses $M_\lambda$, scalar masses $m_{\tilde Q}$, and trilinear couplings $A_{Q_i}$ in terms of derivatives of the field wave-functions~\cite{gr}, \begin{eqnarray} M_\lambda &=&-\frac{1}{2} \left. \frac{\partial \ln S}{\partial \ln \Phi} \right|_0 F_\Phi \\ m_{\tilde Q}^2&=& -\left. \frac{\partial^2 \ln Z_Q}{\partial \ln \Phi \partial \ln \Phi^\dagger} \right|_0 F_\Phi^\dagger F_\Phi \\ A_{Q_i} &=& \left. \frac{\partial \ln Z_{Q_i}}{\partial \ln \Phi} \right|_0 F_\Phi . \end{eqnarray} The symbol ``$|_0$" denotes setting to zero the Grassmann coordinates, $\theta =\bar \theta =0$. Here $S$ and $Z_Q$ are the gauge and matter field wave-functions, with $S$ related to the gauge coupling constant by Re$(S)|_0=g^{-2}/4$. Using Eq.~(\ref{cont}) and $F_\Phi =-m_{3/2}$, see Eq.~(\ref{phiback}), we obtain \begin{eqnarray} M_\lambda &=& -\frac{g^2}{2}\frac{d g^{-2}}{d \ln \mu} m_{3/2} = {\beta_g\over g} m_{3/2} \label{gaugdef} \\ m_{\tilde Q}^2&=&-{1\over 4} \frac{d^2 \ln Z_Q}{d(\ln \mu)^2} m_{3/2}^2= -{1\over 4} \left({\partial\gamma\over\partial g}\beta_g + {\partial\gamma\over\partial y}\beta_y\right)m_{3/2}^2 \label{squarkdef}\\ A_{y}&=&\frac{1}{2} \sum_i \frac{d\ln Z_{Q_i}}{d\ln \mu} m_{3/2}= -{\beta_y\over y} m_{3/2} .\label{adef} \end{eqnarray} Here the sum $\sum_i$ extends over the fields involved in the Yukawa superpotential term with coupling constant $y$, and we have used the renormalization group functions $\gamma(g,y)\equiv {d \ln Z/d\ln\mu}$, $\beta_g(g,y)\equiv {d g/d\ln\mu}$, and $\beta_y(g,y)\equiv {d y/d\ln\mu}$. \subsection{Features of the anomaly-induced soft terms} The soft terms in Eqs.~(\ref{gaugdef})--(\ref{adef}) are determined by the anomalous dimensions of the fields or, in other words, by the violation of the Weyl symmetry in the quantum theory given by the conformal anomaly. Indeed, the supergravity prescription in Eq.~(\ref{cont}) is sufficient to determine the complete form of the soft terms, by means of the technique of ref.~\cite{gr}. The form of the soft terms in Eqs.~(\ref{gaugdef})--(\ref{adef}) is particularly interesting because it is invariant under RG transformations. This means that the analytic continuation into superspace given by Eq.~(\ref{cont}) defines a consistent RG trajectory for the soft terms. The phenomenological appeal of this form of the soft terms resides precisely in this crucial property. In particular, it entails a large degree of predictivity, since all soft terms can be computed from known low-energy SM parameters and a single mass scale, $m_{3/2}$. Also, it leads to robust predictions, since the RG invariance guarantees complete insensitivity of the soft terms from ultraviolet physics. As demonstrated with specific examples in ref.~\cite{giudice}, heavy states do not affect the low-energy parameters in Eqs.~(\ref{gaugdef})--(\ref{adef}), since their effects in the beta-functions and threshold corrections exactly compensate each other. This means that the gaugino mass prediction in Eq.~(\ref{gaugdef}) is valid irrespective of the GUT gauge group in which the SM may or may not be embedded. However, exceptions to ultraviolet insensitivity appear in the presence of gauge singlet superfields~\cite{pr}. The insensitivity from ultraviolet physics not only leads to robust predictivity, but also provides a solution to the supersymmetric flavor problem. Indeed the unknown physics which breaks the flavor symmetry at a high-energy scale $\Lambda_F$ and determines the Yukawa couplings does not leave any visible trace in the anomaly-mediated soft terms. Recall that in gauge mediation the flavor problem is solved by making the soft terms insensitive to any physics above the messenger scale $M$. The parameter $M$ is unknown, and is chosen such that $M<\Lambda_F$. The soft terms vanish above the scale $M$ and therefore their low-energy values are finite and have a logarithmic dependence on $M$. In contrast, in anomaly mediation the soft terms do not vanish at any scale (below the Planck mass $M_{P}$), but their values at low energies are not influenced by physics at any intermediate scale. In order to preserve the attractive properties of the anomaly-mediated soft terms, we have to make sure that other forms of communication of supersymmetry breaking to the observable sector do not give larger contributions. In ordinary gravity mediation, one makes use of tree-level supersymmetry-breaking communication which, in general, dominates over the loop effects of anomaly mediation. If there are no gauge-singlet superfields with scalar vacuum expectation value of order $M_{P}$, then the theory does not contain operators of the form \begin{equation} \int d^2\theta {X\over M_P} {\rm Tr}{\cal W}^\alpha{\cal W}_\alpha + {\rm h.c.}, \end{equation} where $X$ is the Goldstino superfield. Gaugino masses are only generated by higher-dimensional operators and are at best of order $m_{3/2}^{3/2}/M_P^{1/2}$. In particular, this is in general true in theories with dynamical supersymmetry breaking. In this case, the anomaly-mediated effects give the dominant contributions to gaugino masses~\cite{giudice}. It appears at first difficult to forbid or suppress tree-level gravity contributions to scalar masses, which are obtained by couplings in the K\"ahler potential between visible sector fields $Q$ and the Goldstino multiplet $X$, \begin{equation} \label{directcoupling} \int d^4\theta {1\over M_P^2} X^\dagger X Q^\dagger Q. \end{equation} However, the suppression is possible if the K\"ahler potential has the specific structure \begin{equation} K=-3 M_P^2\ln(1-{f_{\rm vis}\over 3 M_P^2} -{f_{\rm hid}\over 3 M_P^2}), \label{kaler} \end{equation} where $f_{\rm vis}$ and $f_{\rm hid}$ are functions of only visible and hidden fields, respectively. This structure could be the result of the underlying fundamental theory such as string theory. However, it is not clear how such a special form of the K\"ahler potential can be stable against radiative corrections. A very interesting possibility, pointed out in ref.~\cite{randall}, is that the supersymmetry-breaking and visible sectors reside on different branes embedded into a higher-dimensional space and separated by a sufficiently large distance. In this case, the structure in Eq.~(\ref{kaler}) is guaranteed by the geometry and not by a symmetry. Thus, all the low-energy soft parameters will arise from anomaly-induced effects. Unfortunately, it turns out that the pure scalar mass-squared anomaly contribution is negative for the sleptons~\cite{randall}. In order to avoid this problem we need to consider other positive soft contributions to the spectrum. This can arise in a number of ways, but any of the solutions will spoil the most attractive feature of anomaly mediation, {\it i.e.} the RG invariance of the soft terms and the consequent ultraviolet insensitivity. This is, in our opinion, the most disappointing aspect of these scenarios. Nevertheless, there are various options to cure this problem without reintroducing the flavor problem. An example is the inclusion of contributions from fields propagating in the bulk space between the two branes~\cite{randall}. Another interesting possibility is a combination of gauge- and anomaly-mediated contributions, discussed in ref.~\cite{pr}. The necessary cure for the slepton masses may completely upset also the mass relations for the other particles (as in the case of the model of ref.~\cite{pr}). However, here we will simply parametrize the new positive contributions to the scalar squared masses with a common mass parameter $m_0$, assuming that the extra terms do not reintroduce the supersymmetric flavor problem. We will see that many of the phenomenological features of an anomaly-induced mass spectrum do not crucially depend on the details of the contributions $m_0$. \subsection{Defining a minimal model} \bigskip In the AMSB scenario, as discussed above, the necessary mass parameters are the gravitino mass, $m_{3/2}$, and the common scalar mass $m_0$, which is required to correct the negative mass-squared of the sleptons. The low-energy soft mass spectrum will be \begin{eqnarray} \label{spectroscopy} M_\lambda &=& {\beta_g\over g} m_{3/2}, \\ m_{\tilde Q}^2&=&-{1\over 4} \left({\partial\gamma\over\partial g}\beta_g + {\partial\gamma\over\partial y}\beta_y\right) m_{3/2}^2 +m_0^2, \label{budda} \\ \label{spectroscopy3} A_{y}&=&-{\beta_y\over y} m_{3/2}. \end{eqnarray} The expressions for the superpartner masses of the minimal particle content and soft parameters are given in the Appendix. We will see that this soft-mass spectrum will give rise to distinctive features which differ from the usual gravity-mediated and gauge-mediated scenarios. Since our working framework is a theory with anomaly-mediated masses and extra universal contributions to the scalar masses, we operationally construct the full supersymmetric spectrum from four input parameters, \begin{equation} m_{3/2}, \, m_0, \, \tan\beta, \, {\rm sign}(\mu ). \end{equation} We treat the $\mu$ and $B_\mu$ masses as derived quantities that combine with other terms in the scalar potential to reproduce correct electroweak symmetry breaking (EWSB). This procedure is done with the one-loop effective potential. Also, we assume that Eq.~(\ref{budda}) is valid at the GUT scale. As previously discussed, the introduction of the scalar mass $m_0$ breaks the RG invariance, and therefore we must define a scale for the boundary condition Eq.~(\ref{budda}). Notice, however, that at the one-loop level with Yukawa couplings neglected, the squark and slepton squared masses are renormalized additively. Therefore, in this case, we do not need to specify at which scale Eq.~(\ref{budda}) is valid. However this is not true, for instance, for the stop and Higgs mass parameters. We find that electroweak symmetry breaking can be accommodated with the above framework. Successful EWSB correlates with values of $|\mu|$ typically between 3 to 6 times the Wino mass as long as $m_0$ is not significantly higher than the anomaly-mediated contributions to the squark masses. Otherwise, $|\mu|$ can be larger. The relative size of $\mu$ with respect to $M_2$ becomes important when considering mass splitting among the degenerate Wino triplet states. This will be considered in more detail in the next section. In Fig.~\ref{spectrum} we demonstrate a subset of superpartner masses using a generically chosen set of input parameters, $m_{3/2} =36\tev$, $\tan\beta =5$, and $\mu < 0$. The choice of $m_{3/2}=36\tev$ determines the gaugino masses to be $M_1=333\gev$, $M_2=119\gev$, and $M_3 =850\gev$. We vary $m_0$ to demonstrate its dependence in the scalar mass spectrum. The squark masses are rather insensitive to values of $m_0$ that raise the slepton masses above their anomaly-mediated tachyonic values. The sleptons, $\tilde e_L$ and $\tilde e_R$, are nearly equal in mass. The extraordinary degeneracy of these slepton masses will be expounded upon in the following section. In Fig.~\ref{spectrum} we also plot the lightest physical Higgs boson mass, $m_h$. This is roughly constant over the range of $m_0$, since this eigenvalue admits only logarithmic sensitivity to supersymmetry breaking scales. Requiring $M_2>90\gev$ and assuming $\tan\beta > 1.8$ (for perturbative unification at the GUT scale), we find a lower bound on the lightest scalar Higgs boson mass of $70\gev$. The lower bound exceeds $100\gev$ for $\tan\beta > 5$. The upper bound on the Higgs boson mass, assuming $M_2<500\gev$ and $m_0<m_{\tilde q}$ is $125\gev$. However, the squark masses are above $3\tev$ when the bound is saturated. Since such high squark masses are not welcome in the loop-corrected Higgs potential, the Higgs mass is expected to be lighter than $125\gev$ in AMSB. On the other hand, the pseudoscalar Higgs mass, $m_A$, depends linearly on the supersymmetry breaking scale, and therefore increases with $m_0$ as shown in the figure. In the next section, we study a few of the unique features of the AMSB spectrum, and how it impacts search capabilities at high-energy colliders. \jfig{spectrum}{spectrum.ps}{ Masses of several states in the supersymmetric spectrum as a function of $m_0$ with $m_{3/2} =36\tev$, $\tan\beta =5$, and $\mu < 0$. The gaugino masses for this choice are $M_1=333\gev$, $M_2=119\gev$, and $M_3 =850\gev$.} \section{Phenomenology} \medskip A unique feature of anomaly-mediated supersymmetry is the gaugino mass hierarchy. To compute the gaugino masses we include next-to-leading corrections coming from $\alpha_s$ and $\alpha_t \equiv y_t^2/4\pi$ two-loop contributions to the beta-functions and weak threshold corrections enhanced by a logarithm. In this approximation, we find \begin{eqnarray} M_1^{NLO}&=& M_1(Q)\left\{ 1+\frac{\alpha}{8\pi \cos^2\theta_W} \left[ -21 \ln \frac{Q^2}{M_1^2} +11 \ln \frac{m_{\tilde q}^2}{M_1^2} +9\ln \frac{m_{\tilde \ell}^2}{M_1^2} \right. \right. \nonumber \\ &+& \left. \left. \ln \frac{\mu^2}{M_1^2} +\frac{2\mu}{M_1}\sin 2\beta \frac{m_A^2}{\mu^2-m_A^2} \ln \frac{\mu^2}{m_A^2}\right] +\frac{2\alpha_s}{3\pi} -\frac{13\alpha_t}{66\pi} \right\} \\ M_1(Q)&=&\frac{11\alpha (Q)}{4\pi \cos^2\theta_W} m_{3/2} \end{eqnarray} \begin{eqnarray} M_2^{NLO}&=& M_2(Q)\left\{ 1+\frac{\alpha}{8\pi \sin^2\theta_W} \left[ -13 \ln \frac{Q^2}{M_2^2} +9 \ln \frac{m_{\tilde q}^2}{M_2^2} +3\ln \frac{m_{\tilde \ell}^2}{M_2^2} \right. \right. \nonumber \\ &+& \left. \left. \ln \frac{\mu^2}{M_2^2} +\frac{2\mu}{M_2}\sin 2\beta \frac{m_A^2}{\mu^2-m_A^2} \ln \frac{\mu^2}{m_A^2}\right] +\frac{6\alpha_s}{\pi} -\frac{3\alpha_t}{2\pi} \right\} \\ M_2(Q)&=&\frac{\alpha (Q)}{4\pi \sin^2\theta_W} m_{3/2} \end{eqnarray} \begin{eqnarray} M_3^{NLO}&=& M_3(Q)\left\{ 1+\frac{3\alpha_s}{4\pi} \left[ \ln \frac{Q^2}{M_3^2} + F\left( \frac{m_{\tilde q}^2}{M_3^2}\right) -\frac{14}{9}\right] +\frac{\alpha_t}{3\pi} \right\} \\ F(x)&=& 1+2x+2x(2-x)\ln x +2(1-x)^2\ln |1-x| \\ M_3(Q)&=&-\frac{3\alpha_s (Q)}{4\pi} m_{3/2} . \end{eqnarray} The higgsino corrections to $M_1$ and $M_2$ are proportional to $\mu / M_{1,2}$ and can become very important in models with large $\mu$, as discussed in ref.~\cite{giudice}. The NLO corrections are significant, especially for $M_2$ where the $6\alpha_s/\pi$ contribution changes the Wino mass by more than $20\%$. The mass ratios of the gauginos $M_1$:$M_2$:$|M_3|$ are approximately $3.3:1:8.8$ at leading order. At NLO, these ratios are changed to $2.8:1:7.1$. This implies that a nearly degenerate triplet of Winos ($\tilde W^\pm$, $\tilde W^0$) are the lightest gauginos. We shall see below that the neutral $\tilde W^0$ is the lightest in the triplet, and is a candidate lightest supersymmetric partner (LSP). In an R-parity conserving theory the $\tilde W^0$ is stable and escapes detection at a high-energy collider. Therefore, visible particles produced in association with the $\tilde W^0$ states will be required to uncover evidence of supersymmetry. It is also possible that the LSP is a sneutrino. This would be the case if the additional contributions to the scalar masses were large enough to generate a positive mass-squared for the sleptons but still smaller than the Wino mass. In this case, Wino decays would generally produce leptons and sneutrinos in the final state. We will consider this possibility in some detail in sect.~\ref{seclep}. \subsection{Mass splitting among Winos} The first step in considering light Wino states is to calculate the mass splitting between the charged and neutral states. For the moment we shall ignore loop corrections and describe the tree-level splitting that develops for light Wino states. Upon integrating out the heavy Bino and Higgsino states, we are left with an effective theory with several operators that could shift the mass of the remaining chargino and neutralino states to be different than $M_2$. Operators of the form ${\cal O}=M_{ab}\tilde W^a\tilde W^b$ will generate mass splittings for the Winos only if $M_{ab}$ transforms non-trivially under $SU(2)$. Because of the symmetry property of the Majorana mass term, $M_{ab}$ must have isospin 2, and the lowest-dimensional operator which generates a mass splitting is \begin{equation} {\cal O}=\frac{1}{\Lambda^3}(H^\dagger \tau^a H)\, (H^\dagger \tau^b H)\, \tilde W^a\tilde W^b , \label{spin} \end{equation} where $\Lambda\sim M_1,\mu$ and $H$ denote the Higgs doublets. Therefore, we see from the above that all mass splittings at tree-level must occur with $m_W^4/\Lambda^3$ suppression. A more detailed formula for the tree-level mass splitting\footnote{To generate a Wino mass splitting, it is also necessary to break the global custodial $SU(2)_V$ defined such that the matrix $\Phi=\pmatrix{H^0_d & H^+_u\cr H^-_d & H^0_u}$, constructed from the two Higgs doublets, transforms as $\Phi \rightarrow V \Phi V^\dagger$ with $V$ unitary. The $\mu$ term is invariant, since it can be written as $\mu \, {\rm det} \Phi$. The symmetry is preserved by electroweak breaking, as long as $\tan \beta =1$, but it is broken by hypercharge effects. Therefore Eq.~(\ref{cazz}) has to vanish in the limit $\tan{\beta}\rightarrow 1$ and $\tan \theta_W \rightarrow 0$.} with $|\mu|\gg M_1,M_2,m_W$ is\footnote{Our sign convention for $\mu$ is set by $W=\mu(H_u^0H_d^0-H_u^+H_d^-)$.} \begin{eqnarray} m_{\tilde \chi^\pm_1}-m_{\tilde \chi^0_1} & = & \frac{m_W^4\sin^22\beta}{(M_1-M_2)\mu^2} \tan^2\theta_W + 2\frac{m_W^4 M_2 \sin 2\beta}{(M_1-M_2)\mu^3 } \tan^2\theta_W \nonumber \\ & &\quad + \frac{m_W^6\sin^3 2\beta}{(M_1-M_2)^2\mu^3} \tan^2\theta_W (\tan^2\theta_W -1) + {\cal O}(\frac{1}{\mu^4}). \label{cazz} \end{eqnarray} When this formula is valid and $\mu$ is determined by the electroweak-breaking condition, the mass splitting is negligible compared to the charged pion mass -- an important mass scale for the phenomenology of Wino decays. In our numerical analysis, we will always calculate the chargino and neutralino mass splittings from the exact formula and not from the expansion in Eq.~(\ref{cazz}), given here only for illustrative purposes. Notice also that, in the large $\tan\beta$ limit, the Wino mass difference becomes \begin{equation} m_{\tilde \chi^\pm_1}-m_{\tilde \chi^0_1} = \frac{M_2 m_W^4}{2\mu^4}\left(1+\frac{2M_2\tan^2\theta_W}{M_1-M_2}\right) + {\cal O}(\frac{1}{\mu^6}),~~~~{\rm for}~\tan\beta \rightarrow \infty . \end{equation} In this limit the mass difference has a further suppression factor, $M_2/\mu$ because the necessary chiral flip cannot originate from the Higgsino mass. The dominant contribution to the Wino mass splitting does not come from the tree-level result described above, but rather due to one-loop corrections in the chargino and neutralino mass matrices. We have done a full numerical calculation of the one-loop corrected chargino and neutralino mass matrices using the formulae of ref.~\cite{pierce}. For the anomaly-mediated spectrum, with only positive mass-squared additional contributions to all the scalar masses, we find that the gauge-boson loop corrections dominate the mass splitting. This is because in the typical anomaly-induced mass spectrum the squark masses are heavy and the $\mu$ parameter is large. Consequently, following the argument that led us to Eq.~(\ref{spin}), we infer that their contribution to the Wino mass splitting is suppressed by $M_W^4/\Lambda^3$. On the other hand, the effect of gauge-boson loops cannot be described by local operators. Isolating this contribution in the limit of large $\mu$, we find (see also refs.~\cite{cheng,feng}) \begin{equation} \label{gaugediff} \Delta_\chi\equiv m_{\tilde\chi_1^\pm} - m_{\tilde\chi_1^0} ={\alpha M_2\over\pi\sin^2\theta_W} \left[f(m_W^2/M_2^2)-\cos^2\theta_W f(m_Z^2/M_2^2)\right] \end{equation} where \begin{equation} f(x)\equiv-{x\over 4}+{x^2\over 8}\ln x+{1\over 2}(1+{x\over 2}) \sqrt{4x-x^2} \left[{\rm arctan}{2-x\over\sqrt{4x-x^2}}-{\rm arctan}{x\over\sqrt{4x-x^2}} \right]. \end{equation} In the limit that $M_2\rightarrow\infty$, the expression in Eq.~(\ref{gaugediff}) simply becomes \begin{equation} \label{loop split} \Delta_\chi= \frac{\alpha\, m_W}{2(1+\cos\theta_W)} \left[1-{3\over 8\cos\theta_W} {m_W^2\over M_2^2}+{\cal O}\left({m_W^3\over M_2^3}\right)\right], \end{equation} which has the asymptotic limit $\Delta_\chi=\alpha\, m_W/[2 (1+\cos\theta_W)] \simeq 165$ MeV. It may appear odd that the mass splitting should asymptote to a constant value as $M_2$ gets arbitrarily massive. This behavior can be understood in momentum space as an infrared mismatch between the self-energies of $\tilde W^+$ and $\tilde W^0$ regulated by $m_W$. Or, equivalently, since $SU(2)$ is a good theory for short distances $r\ll m_W^{-1}$, we can calculate the Coulomb energy of the charged state for large distances $r\gsim m_W^{-1}$ (infrared region) to obtain a mass splitting of approximately $\alpha m_W$. The exact prefactors are given in Eq.~(\ref{loop split}). In Figs.~\ref{split_tb2} and~\ref{split_tb40} we show the total calculated mass splitting as a function of $M_2$ for $\tan\beta =2$ and~10. In our numerical calculation, we include the full one-loop result and we do not use the approximate expressions given in Eq.~(\ref{gaugediff}). The solid curves, from top to bottom, represent $\mu=2M_2$, $\mu=3M_2$, $\mu=5M_2$, and $\mu=\infty$. The dashed curves are the same except that $\mu <0$. The dot-dashed curve is the charged pion mass $m_{\pi^\pm}$. As $\tan\beta$ increases the sign of $\mu$ becomes less and less relevant in the calculation of the mass splitting. When $\tan\beta =40$ the solid and dashed curves are irresolvable. \jfig{split_tb2}{split_tb2.ps}{\protect The mass splitting as a function of $M_2$ for $\tan\beta =2$. The solid curves, from top to bottom, represent $\mu=2M_2$, $\mu=3M_2$, $\mu=5M_2$, and $\mu=\infty$. The dashed curves are the same except for the opposite sign of $\mu$. The dot-dashed curve is the charged pion mass $m_{\pi^\pm}$.} \jfig{split_tb40}{split_tb10.ps}{\protect The mass splitting as a function of $M_2$ for $\tan\beta =10$. The solid curves, from top to bottom, represent $\mu=2M_2$, $\mu=3M_2$, $\mu=5M_2$, and $\mu=\infty$. The dashed curves are the same except for the opposite sign of $\mu$. The dot-dashed curve is the charged pion mass $m_{\pi^\pm}$.} In an anomaly-mediated spectrum with radiative electroweak symmetry breaking, the typical relation between $M_2$ and $\mu$ is $3\lsim |\mu|/M_2\lsim 6$. This is true as long as the squark masses are not increased significantly beyond their anomaly-mediated baseline values from the universal mass contributions that lift the slepton mass-squared to positive values. This is also acceptable from naive fine-tuning arguments. Larger values of $\mu$ lead to an unnatural Higgs potential. For $|\mu|/M_2= 5$ we find from Figs.~\ref{split_tb2} and~\ref{split_tb40} that the mass splitting is significantly above $m_{\pi^\pm}$ such that $\tilde W^\pm \rightarrow \tilde W^0\pi^\pm$ is kinematically allowed and is the dominant decay mode. This remark is also true even for extraordinarily large values of $\mu$ as long as $M_2\gsim 80\gev$. \subsection{Finding supersymmetry with dileptons} The precise calculation of the mass splitting is crucial since in ref.~\cite{gunion} it was demonstrated that if $m_{\pi^\pm}\lsim m_{\tilde \chi^\pm_1}-m_{\tilde \chi^0_1} \lsim 1\gev$ then the $\tilde W^\pm$ will decay too fast to use a quasi-stable charged particle analysis, with dedicated triggers. However, the decays are not prompt, and so analyses of events triggered by other means could see a stiff charged particle track that subsequently terminates in the vertex detector. The difficulty is triggering the event. One way to trigger such events is to produce the Winos in associated production with a standard model particle, such as a gluon at hadron colliders or a photon at $e^+e^-$ colliders. Triggering on high-$p_T$ monojets or high-energy photons at these colliders then may be an effective way to trigger the events and save them for future analysis~\cite{gunion,thomas,feng}. At the analysis stage a kink in the vertex detector, or a terminating stiff track, would then indicate a non-SM underlying process. Here we pursue another direction for discovery. We can utilize production and subsequent decays of other SUSY particles as a way to trigger on the events and learn more about the theory. For example, if sleptons or squarks are produced in a hadronic collision, they will cascade decay to high-$p_T$ SM particles and charged and/or neutral Winos. The SM particles can be used for the trigger, and the cascade decays can be used to learn something about the spectroscopy of the theory. Our example process is left-handed slepton and sneutrino production at the Tevatron which cascades into $l^\pm l^\mp+X_{D}$, where $X_D$ is a displaced vertex from one or two $\tilde W^\pm \rightarrow \tilde W^0\pi^\pm$ decays. These displaced vertices are heavy charged particle tracks which stop in the vertex detector and produce very soft pions that may or may not be detectable. The dilepton events are produced through \begin{eqnarray} p\bar p \rightarrow \tilde \nu_L\tilde l^{\pm}_L & \rightarrow & l^\pm l^\mp \tilde W^\pm\tilde W^0 \rightarrow l^\pm l^\mp +X_D+E^{\rm miss}_T \\ p\bar p \rightarrow \tilde \nu_L\tilde \nu_L & \rightarrow & l^\pm l^\mp \tilde W^\pm\tilde W^\mp \rightarrow l^\pm l^\mp +X_D+E^{\rm miss}_T . \end{eqnarray} In Fig.~\ref{likesign} we plot the total cross-section of such events for one flavor ($\mu^\pm\mu^\mp+X_D$) at the $\sqrt{s}=2\tev$ Tevatron. We require the pseudo-rapidity to be within $|\eta|<2$ for both leptons, and we require the leading lepton to have $p_T>10\gev$, and the next lepton to have $p_T>5\gev$. The total rate presented in Fig.~\ref{likesign} is calculated at leading order. We have included a total of 50\% suppression of the naive LO result from jet veto and lepton identification efficiency. Our conclusion based on Fig.~\ref{likesign} is that left-handed sleptons with mass less than $200\gev$ would be discovered at the Tevatron if $M_2\lsim m_{\tilde \nu_L}-10\gev$ (for leptons to have high enough $p_T$ for triggering) and if the Tevatron reaches at least $30\xfb^{-1}$ integrated luminosity. This result is based on the requirement that more than 10 $l^\pm l^\mp+X_D$ events will occur for each flavor. We conservatively choose a 10 event requirement in order to ensure that our mass reach conclusion will remain valid if the dilepton identification efficiency were to be as low as $50\%$ for the $p_T$ acceptance cuts given above. Other modes such as $\mu^\pm +X_D$ are possible in $\tilde \nu_L$ and $\tilde \mu_L$ production, and could confirm and extend the mass reach capabilities of the dilepton mode. \sfig{likesign}{likesign.ps}{Dilepton signal from left-handed smuon and sneutrino production at the Fermilab Tevatron with $2\tev$ center of mass energy and $M_2=90\gev$. Acceptance cuts of the leptons are described in the text. The different curves are for $\tan\beta =1$, which makes $m_{\tilde \mu_L}=m_{\tilde \nu_L}$, and for $\tan\beta = \infty$ which maximizes the hypercharge $D$-term splitting such that $m_{\tilde \mu_L}^2=m_{\tilde \nu_L}^2+m_W^2$. With $30\xfb^{-1}$ the Tevatron will record more than 10 such events for each lepton flavor if $m_{\tilde \nu_L}<200\gev$.} The dilepton signal discussed above is special since it is essentially background free with the displaced vertices present. The possibility of prompt Wino decays arises if $\mu$ is sufficiently light to yield a chargino/neutralino mass splitting above $1\gev$. This would make the charged Wino decay promptly into a neutral Wino plus other states too soft to admit into the event description. Also, if the top squarks are reduced from additional negative scalar mass sources, the mass splitting could be greatly enhanced by loop corrections involving third family sfermions and fermions. In these cases, special triggers or analyses based on decay kinks of the charged Wino could not be relied upon, but the dilepton signal would remain useful. \subsection{Degeneracy of sleptons} Another striking feature of the anomaly-mediated model with additional universal scalar terms is the near degeneracy of the left and right sleptons of the first two generations. The mass-squared splitting is somewhat insensitive to $m_0$, \begin{eqnarray} \label{ersplit} \Delta_{\tilde e} & = & m^2_{\tilde e_L}-m^2_{\tilde e_R} = (11\tan^4\theta_W-1)\frac{3}{2}M_2^2 + \left(-\frac{1}{2}+2\sin^2\theta_W\right) m^2_Z\cos 2\beta \nonumber \\ & & \qquad\qquad\quad\quad\quad +\frac{1}{8\pi^2}\left( \frac{9}{5}g_1^2 M^2_1 - 3g_2^2 M_2^2\right) \ln \frac{m_{\tilde e_R}}{m_Z} \nonumber \\ & \simeq & 0.037 \left( -m_Z^2\cos 2\beta+M_2^2 \ln \frac{m_{\tilde e_R}}{m_Z} \right). \end{eqnarray} The first term is the tree level anomaly-induced splitting, the second term is the hypercharge $D$-term splitting induced by electroweak breaking, and the third term is the one loop, leading log mass splitting induced by renormalizing the masses to their own scale. It is a numerical accident that the value of $\sin^2\theta_W$ is such that the $M_2^2$ coefficient in the first term of Eq.~(\ref{ersplit}) is nearly zero. If, \begin{equation} \label{sin squared} \sin^2\theta_W=\frac{1}{1+\sqrt{11}}=0.2317 \end{equation} then the tree-level coefficient of $M_2^2$ would be identically zero. The actual value of $\sin^2\theta_W(m_Z)$ is $0.2312\pm 0.0003$~\cite{PDG} in the $\overline {MS}$ scheme and is extraordinarily close to the value in Eq.~(\ref{sin squared}) required to make the $M_2^2$ coefficient in the tree-level mass splitting vanish. It is also a numerical accident that the hypercharge $D$-term coefficient is suppressed since $\sin^2\theta_W\simeq 1/4$. Although the coefficient is not as spectacularly suppressed as the $m_{3/2}^2$ coefficient, it is multiplied by a fixed scale $m_Z^2$. Therefore, for a given value of $\tan\beta$ the mass squared difference remains constant regardless of how heavy the sleptons may be. The degeneracy of the slepton can be characterized by the fractional difference, \begin{equation} \frac{m_{\tilde e_L}-m_{\tilde e_R}}{m_{\tilde e_R}} = -1 +\sqrt{ 1+\frac{\Delta_{\tilde e}}{m^2_{\tilde e_R}}} ~\simeq \frac{1}{2}\frac{\Delta_{\tilde e}}{m^2_{\tilde e_R}} . \end{equation} In Fig.~\ref{diff} we plot contours of the relative mass splitting in the $M_2$-$m_{\tilde e_R}$ plane. The mass splitting is less than a few percent over most of parameter space. It exceeds 5\% only when $M_2>350\gev$. However, the squark masses in this case are over $2\tev$, which induces a considerable fine-tuning in the one-loop Higgs potential. Therefore, it is not expected that $M_2$ is so high, implying that the slepton mass differences should be no more than a few percent over the relevant parameter space. \jfig{diff}{diff.ps}{Contours of $100\% \times (m_{\tilde e_L}-m_{\tilde e_R})/m_{\tilde e_R}$ in the $M_2$-$m_{\tilde e_R}$ mass plane with large $\tan\beta$, which maximizes the mass splitting.} The resolution of slepton masses from end-point lepton distributions of the slepton decays is approximately $2\%$ at an $e^+e^-$ or a muon collider~\cite{NLC report}. Note that at a polarized linear collider it will not be difficult to determine that both left and right sleptons are being produced even if they are degenerate. This can be accomplished most effectively by comparing total rate of sleptons production with the asymmetry of production for polarized beams~\cite{NLC report}. Degenerate sleptons are not expected in the usual supergravity or gauge-mediated scenarios, where $m_{\tilde e_R}$ is generally lighter than $m_{\tilde e_L}$. An exception to this is in the minimal supergravity model with $m_0\gg m_{1/2}$. However, given the current limits on $m_{1/2}$ from gaugino searches at the Tevatron and LEP, degenerate sleptons will only occur at very high mass. The above discussion is based on the assumption that the additional contributions to the slepton masses are universal. Since the anomaly-mediated mass differences is accidentally negligible, the degeneracy of the sleptons becomes a test of the additional mass contributions. Other approaches to the slepton tachyonic problem do not necessarily imply degenerate slepton mass eigenstates~\cite{pr}. \subsection{LEP2 Signals} \label{seclep} \bigskip At LEP2 many signatures are possible in the AMSB scenario. The charged Winos have a large production cross-section as long as they are kinematically accessible and as long as the sneutrino $t$-channel amplitude does not significantly interfere destructively with gauge boson $s$-channel amplitudes. Production of charginos at LEP has been the topic of many studies in supersymmetry phenomenology at LEP~\cite{leprep}. However, most of these studies have assumed that the LSP, the lightest neutralino, is more than a few GeV below the lightest chargino mass. In AMSB this is no longer the case. We expect that the lightest neutralino and charginos form a nearly degenerate $SU(2)$ triplet, as discussed in previous sections. Since the chargino is only slightly above the neutralino in mass, the process $e^+e^-\rightarrow \tilde W^+\tilde W^-$ will be accompanied by very soft visible final states from decays such as $\tilde W^+\rightarrow \pi^+\tilde W^0$. These soft final states cannot be triggered on, which has led others~\cite{gunion,thomas,feng} to suggest triggering on initial state photon radiation and then searching for soft, displaced tracks at the analysis level. However, there are other potential signatures of supersymmetry when Winos are the lightest gauginos. To enumerate them we must consider the other states in the supersymmetric spectrum which may be produced in collisions or as decay products of Wino production. In the AMSB scenario, the degenerate left and right sleptons and the sneutrino are the most important states at LEP after the gauginos. The ratio of their masses to the Wino masses is unknown in our framework, but it is more natural that they be somewhat light in order to keep the Higgs scalar potential from being fine-tuned. Therefore, considering phenomenological implications of light sleptons at LEP is useful. There are many permutations to the relative ordering of $M_2$, $m_{\tilde \nu_L}$, and $m_{\tilde e}\equiv m_{\tilde e_{L,R}}$. Recall that the relationship between $m_{\tilde e}$ and $m_{\tilde\nu_L}$ is \begin{equation} m^2_{\tilde e_L}=m^2_{\tilde\nu_L}-m^2_W\cos 2\beta. \end{equation} We can provide the general phenomenological features using a graph in the $M_2$-$m_{\tilde\nu_L}$ plane for LEP2 running at $\sqrt{s}=200\gev$. The results of the present and future experimental analyses combining the searches at LEP2 in different channels will be best presented as exclusion or discovery regions in the $M_2$-$m_{\tilde\nu_L}$ plane. The LEP1 limit of Wino masses is slightly above $m_Z/2$, and the limit on sneutrino masses is slightly below $m_Z/2$; therefore, we begin the axes of Fig.~\ref{m2msnu} at $m_Z/2$. The lines represent kinematic boundaries. For example, the top dashed line is where $m_{\tilde e}=\sqrt{s}/2$, and for all $m_{\tilde\nu_L}$ values above that line, $\tilde e\tilde e$ production is not possible. The precise locations of the dashed lines depend on the choice of $\tan\beta$, which we choose to be $\tan\beta =3$ for this figure. \bfig{m2msnu}{m2msnu.ps}{Signatures of the AMSB scenario at LEP2. Each blocked region has a unique mass hierarchy among $\tilde \nu_L$, $\tilde e_L$ and $M_2$, and therefore leads to different signatures which are contained within the parenthesis. The meaning of ``$(\gamma )$'', for example, is $e^+e^-\rightarrow \gamma \tilde W^+\tilde W^-\rightarrow \gamma +\mE$. We assume, as is justified by EWSB analysis, that $M_2$ is very close to the mass of the nearly degenerate lightest charginos and neutralino. The mass splitting between $m_{\tilde \nu_L}$ and $m_{\tilde e}$ is due to a hypercharge $D$ term, and its value is calculated with $\tan\beta =3$ for this figure.} Each region in the figure constitutes a different ordering of the mass eigenstates, and will produce a different set of useful observables with which to probe the theory. These observables are given inside the parentheses. For example, in the small triangular region surrounding the point $(M_2,m_{\tilde\nu_L})=(95\gev ,60\gev )$ one can search for $\gamma +\mE$, $l^+l^-+\mE$, and up to four electrons plus missing energy. Specific processes which lead to these signatures include, \begin{eqnarray} e^+e^- & \rightarrow & \gamma \tilde \nu_L\tilde\nu_L \rightarrow \gamma +\mE \\ e^+e^- & \rightarrow & (\tilde W^+ \tilde W^-~{\rm or}~ \tilde l^-\tilde l^+)\rightarrow l^+l^-+\mE \\ e^+e^- & \rightarrow & \tilde W^0 \tilde W^0\rightarrow \tilde l^+ l^-\tilde l'^+ l'^-\rightarrow l^+l^- l'^+l'^- +\mE. \end{eqnarray} Some of the leptons may be softer than others because of reduced phase space in a decay of a massive sparticle into a lepton and a sparticle with mass near its parent. Near the boundaries of the curves, it is often the case that some leptons are not energetic enough, and care must be taken in the analysis to identify them. Finally, the stau sleptons may be lighter than the other sleptons, leading to more $\tau$ lepton final states than other leptons. Although efficiency in identifying $\tau$ leptons is smaller than the others, it is possible at large $\tan\beta$ to have large mixing among the $\tilde \tau_L$ and $\tilde \tau_R$ sleptons to produce a mass eigenstate accessible to LEP whereas the other sleptons are not. \section{Gravitino cosmology} \medskip A distinctive feature of the AMSB scenario is that the gravitino is much heavier than the supersymmetric partners of ordinary particles. The reason for this is that the AMSB masses are suppressed by a loop factor relative to the gravitino mass $m_{3/2}$. In particular, one finds that the gravitino--Wino mass ratio is $m_{3/2}/M_2 \simeq 300$. A large gravitino mass is cosmologically advantageous for solving the gravitino problem \cite{weinberg}. This problem occurs when gravitino decay products disrupt the light element abundance during nucleosynthesis. Even a period of inflation is not sufficient to solve this problem since gravitinos are thermally produced during the reheating phase of the universe~\cite{ellis}. Thus, in order for the gravitino decay products to be harmless during nucleosynthesis one either requires that the gravitino decays before affecting nucleosynthesis or that the reheating temperature of the universe be bounded from above. The gravitino number density in units of the photon number density after the inflationary epoch is~\cite{moroi} \begin{equation} \frac{n_{3/2}}{n_\gamma}(T)= 2\times 10^{-9} g_*(T)~ T_{13}( 1-0.03 \ln T_{13}) . \end{equation} Here $T_{13}$ is the reheating temperature after inflation in units of $10^{13}\gev$ ({\it i.e.}, $T_{13}\equiv T_R/10^{13}$ GeV), and $g_*(T)$ counts the massless degrees of freedom at the temperature $T$ including a factor of $7/8$ for fermions and the dilution factor for frozen-out species. The gravitino decay width is \begin{equation} \Gamma_{3/2} = {1\over 4}\left(N_g+{N_m\over 12}\right) {m_{3/2}^3\over M_P^2} \simeq 5.1 \left({m_{3/2}\over 50~{\rm TeV}}\right)^3 {\rm sec}^{-1}. \end{equation} Here $M_P=1.2\times 10^{19}$ GeV, $N_g$ and $N_m$ are the number of gauge and matter decay channels, and we have summed over all the SM particle content ($N_g$=12, $N_m$=49). Immediately after the gravitino decays, the temperature of the universe is given by \begin{equation} T_D=\left( \frac{45\, \Gamma_{3/2}^2M_P^2}{4\pi^3g_*(T_D)}\right)^{1/4}= 2.7 \left( \frac{10.75}{g_*(T_D)}\right)^{1/4} \left({m_{3/2}\over 50~{\rm TeV}}\right)^{3/2} {\rm MeV}. \end{equation} Therefore for $m_{3/2}\lsim 60$ TeV a detailed analysis of effects of gravitino decay during nucleosynthesis is necessary. The particular gravitino decay products that cause the main interference during nucleosynthesis at early times ($\sim 1$ second) are hadronic showers~\cite{reno}. Photodissociations are not relevant at early stages of the nucleosynthesis epoch since the destructive photon-nucleus interactions are much less probable than photon-photon interactions. The overall effect of hadronic decay products is to convert protons into neutrons, and consequently the $^4$He abundance is increased since the additional neutrons that are produced are synthesised into $^4$He. Thus, using the observational upper limit on the primordial helium abundance $Y(^4{\rm He})<0.25$, we obtain an upper bound on the reheat temperature~\cite{sarkar} which is depicted in Fig.~\ref{reheat}. \jfig{reheat}{reheat.ps}{The upper bound on the reheat temperature as a function of the gravitino mass $m_{3/2}$ and the corresponding Wino mass $M_2$. The observational limit upper limit on the primordial helium abundance is $Y(^4{\rm He})<0.25$. Therefore, the region above the line is excluded, and an upper bound on $T_R$ results.} For $m_{3/2}\lsim 40$ TeV, the typical bound on the reheat temperature is $T_R\sim 10^{9}$ GeV. This is typically less constraining than in the usual gravity-mediated scenarios with weak-scale gravitino mass~\cite{sarkar}. As the gravitino mass increases, the upper bound on the reheat temperature becomes less significant, and completely evaporates for $m_{3/2}\gsim 60$ TeV, since the gravitinos decay well before the start of nucleosynthesis. For $m_{3/2}\gsim 60$ TeV, we may be concerned that the entropy produced by the gravitino decay excessively dilutes the baryon--to--photon ratio obtained by a primordial baryogenesis mechanism. However, this is never the case. Actually for \begin{equation} {m_{3/2}\over 50~{\rm TeV}} >8\times 10^{-4} \left(\frac{g_*(T_D)} {10.75}\right)^{1/2} T_{13}^2( 1-0.03\ln T_{13})^2 , \end{equation} the gravitino decays before dominating the universe and the entropy release is not dangerous. We have checked that, even when the gravitino matter-dominates the universe, the entropy production is not problematic. Therefore we conclude that when $m_{3/2}\gsim 60$ TeV, there is no upper bound on the reheat temperature arising from nucleosynthesis. However, bounds on the reheating temperature when $m_{3/2}\gsim 60\gev$ do come from considerations of the Wino energy density. The Wino thermal relic abundance $\Omega_{LSP}^{TH}$ does not play a significant cosmological role. Indeed Wino annihilations into gauge bosons in the early Universe are very efficient and lead to~\cite{giudice} \begin{equation} \Omega_{LSP}^{TH} h^2 \simeq 5\times 10^{-4} \left( \frac{M_2}{100~{\rm GeV}} \right)^2 . \end{equation} On the other hand, a non-thermal production of LSPs is generated by the gravitino decay~\cite{frie}. Since this decay occurs below the Wino freeze-out temperature, the LSP abundance is easily determined by assuming that each decaying gravitino produces a single Wino. The LSP is relativistic at decay time and becomes non-relativistic at a typical temperature $T\sim T_D M_2/m_{3/2}$, after red-shifting. The predicted relic abundance is \begin{equation} \Omega_{LSP}^{G} h^2\simeq 30 \left(\frac{M_2}{100\, {\rm GeV}}\right) T_{13}(1-0.03\ln T_{13}). \end{equation} The requirement for not overclosing the universe leads to a bound on the reheat temperature $T_R\lsim 10^{11}\gev$. Of course, if R-parity is not conserved then this bound from LSP relic abundance is no longer relevant. In this case, for gravitino masses $m_{3/2}\gsim 60$ TeV, one can then contemplate using leptogenesis or even GUT baryogenesis mechanisms to generate the baryon asymmetry of the universe, consistently with the AMSB gravitino cosmology. On the other hand, when $T_R\simeq 10^{10}$--$10^{11}\gev$ the relic abundance of LSPs from gravitino decays is near critical density, providing a natural source of dark matter. Finally, we comment on another positive aspect of the heavy gravitino cosmology. We can avoid the cosmological Polonyi problem that arises in the usual gravity-mediated scenario when the gauge singlet Polonyi field acquires a Planck scale vacuum expectation value but decays relatively late. In the AMSB scenario there is simply no need for the Polonyi field since the gaugino masses arise from a quantum anomaly. \section{Conclusion} \bigskip In summary, anomaly-induced masses are always present when supersymmetry is broken. When these AMSB contributions dominate and yield all the gaugino masses as well as adding to universal scalar masses, a unique spectrum results which has important differences from other models of supersymmetry. Several of these unorthodox features that arise in low-energy supersymmetry from the AMSB scenario include, \begin{itemize} \item The ratio of gaugino masses $M_1:M_2:|M_3|$ is approximately $2.8:1:7.1$ when loop corrections are included, and $M_2\simeq m_Z$; \item The lightest supersymmetric particle is most often the Wino, but may also be the sneutrino; \item Dilepton signals with displaced vertices are are useful signals for this scenario at LEP and Tevatron; \item The anomaly-induced contribution to left and right slepton masses is accidentally degenerate. This remains true if the required, additional sources for slepton masses are universal; \item LEP signatures are sensitive to the hierarchy of sneutrino, slepton and Wino masses. The searches in the different channels can be simply combined to give exclusion plots in the $M_2$-$m_{\tilde\nu_L}$ plane; \item The gravitino mass is much heavier than the masses of the other sparticles. Consequently, the cosmological problem associated with gravitino decays during nucleosynthesis is alleviated over much of parameter space; \item In spite of its negligible thermal relic abundance, neutral Winos can form the galactic dark matter, since they are copiously produced, below their freeze-out temperature, from the primordial gravitino decays. \end{itemize} Discovery of several of the above phenomenological implications is necessary to gain confidence that the AMSB scenario is a proper description of nature. \section*{Appendix} Using Eqs.~(\ref{spectroscopy})-(\ref{spectroscopy3}), the anomaly-mediated spectrum is \begin{eqnarray} \wompit{M_1} & = & \frac{33}{5} \frac{g_1^2}{16\pi^2} m_{3/2} \\ \wompit{M_2} & = & \frac{g_2^2}{16\pi^2} {m_{3/2}}\\ \wompit{M_3} & = & -3\frac{g^2_3}{16\pi^2}{m_{3/2}}\\ \womp{m^2_{\tilde t_R}} & = & \left( -\frac{88}{25}g^4_1+8g_3^4 +2y_t\hat\beta_{y_t} \right) \frac{m_{3/2}^2}{(16\pi^2)^2} \\ \womp{m^2_{\tilde b_R}} & = & \left( -\frac{22}{25}g^4_1+8g_3^4+2y_b\hat\beta_{y_b} \right) \frac{m_{3/2}^2}{(16\pi^2)^2}\\ \womp{m^2_{\tilde Q_3}} & = & \left( -\frac{11}{50}g^4_1-\frac{3}{2}g_2^4+8g^4_3 +y_t\hat\beta_{y_t}+y_b\hat\beta_{y_b} \right) \frac{m_{3/2}^2}{(16\pi^2)^2} \\ \womp{m^2_{H_u}} & = & \left( -\frac{99}{50} g^4_1 -\frac{3}{2}g^4_2 +3y_t\hat\beta_{y_t} \right) \frac{m_{3/2}^2}{(16\pi^2)^2} \\ \womp{m^2_{H_d}} & = & \left( -\frac{99}{50}g^4_1 -\frac{3}{2}g^4_2 +3y_b\hat\beta_{y_b} + y_{\tau}\hat\beta_{y_\tau} \right) \frac{m_{3/2}^2}{(16\pi^2)^2} \\ \womp{m^2_{\tilde L_3}} & = & \left( -\frac{99}{50}g^4_1-\frac{3}{2}g_2^4 +y_{\tau}\hat\beta_{y_\tau}\right) \frac{m_{3/2}^2}{(16\pi^2)^2} \\ \womp{m^2_{\tilde \tau_R}} & = & \left( -\frac{198}{25}g^4_1 +2y_\tau \hat\beta_{y_\tau}\right) \frac{m_{3/2}^2}{(16\pi^2)^2} \\ A_{y_t} & = & -\frac{\hat\beta_{y_t}}{y_t}\frac{m_{3/2}}{16\pi^2} \\ A_{y_b} & = & -\frac{\hat\beta_{y_b}}{y_b}\frac{m_{3/2}}{16\pi^2} \\ A_{y_\tau} & = & -\frac{\hat\beta_{y_\tau}}{y_\tau} \frac{m_{3/2}}{16\pi^2} . \end{eqnarray} where \begin{eqnarray} \hat\beta_{y_t} & = & 16\pi^2\beta_{y_t} =y_t\left( -\frac{13}{15}g^2_1 -3g_2^2 -\frac{16}{3}g^2_3 +6y_t^2 +y_b^2 \right) \\ \hat\beta_{y_b} & = & 16\pi^2\beta_{y_b} =y_b\left( -\frac{7}{15}g^2_1-3g^2_2 -\frac{16}{3}g^2_3 +y_t^2 +6y_b^2+y_\tau^2\right) \\ \hat\beta_{y_\tau} & = & 16\pi^2\beta_{y_\tau} =y_\tau \left( -\frac{9}{5}g^2_1-3g^2_2+3y_b^2 +4y_\tau^2 \right). \end{eqnarray} The first two generation squark and slepton masses are obtained by appropriately changing the Yukawa couplings to first and second generation Yukawa couplings.
\section{Introduction} \label{sect1} Asymptotically free models gained importance both in High Energy and in Condensed Matter Physics. This is obvious for the latter, where the low dimensional phenomena invoke effective theories below the upper critical dimension, $d=4$, which are super-renormalizable and their Hamiltonians contain asymptotically free coupling constants only. In the former case the the asymptotically free models can provide structureless high energy physics where the cutoff can be pushed away at will. What kind of non-perturbative mechanisms do we encounter as we follow the renormalization group flow of an asymptotically free model? To distinguish different finite energy mechanisms we recall that unless the model possesses unbroken scale invariance there are always two scaling regimes, an ultraviolet and an infrared one separated by a crossover at the characteristic scale of the model. If there is a dimensional coupling constant, say a mass parameter $m$, in the Lagrangian then the UV and IR scaling regimes are separated by a crossover at $p=p_{cr}\approx m$ because $m^2/p^2$ or $p^2/m^2$ is treated as a small quantity in the UV or the IR side, respectively. The IR scaling laws are trivial because the radiative corrections to the evolution are suppressed by $p^2/m^2$ and what is left is the scale dependence governed by the canonical dimensions of the coupling constants. If there are no dimensional coupling constants in the Lagrangian, i.e. the model possesses classical scale invariance then a crossover can be identified where one of the coupling constants reach the value 1, $g(p_{cr})=1$. The expression for the dimensionless running coupling constant $g(p)$ must contain a dimensional parameter, usually called the $\Lambda$-parameter and $p_{cr}\approx\Lambda_{\phi^4}$.\footnote{It is worthwhile noting that one can introduce the $\Lambda$-parameter even in the presence of any parameter with positive mass dimension in the Lagrangian so long as the IR dynamics remains stable when this latter is removed. In fact, let us write this parameter as $m^\ell$ where $m$ defined in this manner is the mass scale of the model. The running coupling constant depends on the ratio $m/p$, $g(p)=f(m/p)$ where the limit $m\to0$ is convergent. In other words, the evolution of the classically scale invariant system is recovered as $p\to\infty$.} Non-perturbative effects may originate in the IR scaling regimes. The classically scale invariant models at the lower critical dimension, such as the two-dimensional sigma and the Gross-Neveu model do not support long range order. Thus the IR scaling laws must contain a dynamically generated mass scale, an effect generated by the infrared or collinear divergences. The asymptotically free coupling constants may generate non-perturbative effects in the UV scaling regime, as well, due to their growth. The question is whether the scale parameter $\Lambda_{\phi^4}$ or the mass $m$ is reached first as we lower the cutoff. The system remains perturbative for $m>\Lambda_{\phi^4}$ because the IR scaling laws cut off the growth of the asymptotically free coupling constants before they would reach a dangerously large value. The dynamics of the modes $p<\Lambda_{\phi^4}$ becomes non-perturbative when $\Lambda_{\phi^4}>m$. The four-dimensional Yang-Mills models develop linearly rising potential between static external charges. This is believed to happen due to the large value of the asymptotically free coupling constant at the ultraviolet side of the crossover, indicated by the infrared Landau pole in perturbative QCD. The quark masses are supposed to play no important role in the vacuum structure and the chiral limit $m\to0$ is assumed to be safe and convergent, though non-perturbative. The infrared singularities are easier to isolate than the non-perturbative effects arising at the IR end of the UV scaling regime. It has already been achieved for non-asymptotically free models \cite{irqed} and superrenormalizable theories \cite{jate}. The more complex non-perturbative features should arise at the ultraviolet side of the crossover where no asymptotic analysis is available. The goal of the present paper is the study of the scaling laws and their consequences at the IR end of the UV scaling regime. According to our best knowledge this source of the non-perturbative effects in asymptotically free theories has not been studied before in a systematic manner. The problem is rather involved and demanding, we restrict ourselves to outline only the way the more detailed analysis should be done. In order to defuse the infrared problem we turn to the scalar model with polynomial interaction in the phase where no spontaneous symmetry breaking occurs. The term $\phi^n$ in the Lagrangian of the $d$-dimensional scalar model is asymptotically free for $n<2d/(d-2)$, so we have to trace the evolution of a large number of coupling constants at low dimensions. We present numerical evidences that the polinomial interactions at the lower critical dimension or below where infinitely many operators are relevant in the UV scaling regime always become strong, non-perturbative. The study of the asymptotically free three-dimensional $\phi^6$ model can bring some light into the IR Landau pole problem of four-dimensional gauge theories by considering the role of the non-renormalizable couplings in the IR extrapolation of the UV scaling regime. We will show that they affect considerably the evolution of the relevant couplings. This points to the possibility that the non-renormalizable couplings could suppress the Landau pole. The full solution of the theory where all possible coupling constants are followed by the renormalization group equation should necessarily yield a non-singular renormalization group flow so long as the theory possesses local interactions. This suggests that the introduction of the hadronic composite operators in QCD could defuse its Landau pole problem \cite{polonyi95}. Another issue addressed in this paper is the manner in which the influence of the non-renormalizable operators on the dynamics is suppressed and the universal physics is reached as the cutoff is lowered. The usual argument based on the linearization of the blocking relation is not obviously applicable due to the presence of infinitely many relevant operators. This question is discussed in the framework of lattice regulation by tailoring the Wegner-Houghton equation for lattice regulated models. As a result, the approach to the low energy physics can be compared for the momentum space cutoff regularization in the continuum and for the lattice regularization, and the universality, the regulator independence, can be established well beyond the linearized approximation of the blocking relations. The organization of this paper is the following. The infinitesimal renormalization group step, the Wegner-Houghton equation, is described in Section II. The running coupling constants are introduced and their evolution equations are given in Section III for scalar models. The asymptotic UV evolution is discussed in Section IV for $d=2,~3$ and 4. Section V is devoted to the demonstration of the difficulties in finding a perturbative model in $d=2$. The Wegner-Houghton equation is derived in the lattice regularization and its solution is presented in Section VI. Section VII is for the conclusions. \section{The Wegner-Houghton equation} There are different ways the mixing of a large number of operators can be traced down. The Wegner-Houghton equation \cite{wh73}, which we use in the local potential approximation in this work, is the simplest implementation of the Kadanoff-Wilson blocking \cite{wrg} in the momentum space and produces the cutoff dependence of the bare action along the renormalized trajectory. Other methods work with the effective action where the infrared cutoff dependence is sought \cite{others}. Different schemes should agree in the infrared limit where few long wavelength modes are left only in the system. We shall make two approximations in computing the blocked action, the truncation of the gradient expansion at the leading order, the local potential approximation \cite{locp}, and the truncation of the Taylor expansion of the local potential in the field variable \cite{trpot}. The higher order terms of the gradient expansion are non-renormalizable according to the power counting. We believe that these coupling constants which are irrelevant in the ultraviolet scaling regime do not modify our qualitative conclusion. The Wegner-Houghton (WH) equation~\cite{wh73} describes the evolution of the effective action as the cutoff is lowered. As mentioned in the Introduction we shall consider a scalar model with an intrinsic mass scale which allows to clearly distinguish the UV and IR scaling regimes. We derive the WH equation for a scalar field theory by using a sharp momentum space cutoff~\cite{polonyi95,polonyi97}: we will call this regularization procedure the \textit{continuum regularization.} In Section~\ref{lattice} we consider an alternative, the \textit{lattice regularization.} Denote the bare action by $S_k[\phi]$, where $k$ is the UV cutoff. Then, according to the usual Wilson-Kadanoff procedure, \begin{equation} e^{-\frac{1}{\hbar}S_{k'}[\phi]}=\int\mathcal{D}\phi'\, e^{-\frac{1}{\hbar}S_{k}[\phi+\phi']} \label{blocking} \end{equation} where $k'<k$ in the Euclidean space-time. The Fourier transform of the fields $\phi(x)$ and $\phi'(x)$ is non-vanishing only for $p<k'$ and $k'<p<k$, respectively. The right hand side is evaluated by means of the loop expansion, so that Eq.~(\ref{blocking}) gives \begin{equation} S_{k'}[\phi]=S_k[\phi+\phi'_0]+\frac{\hbar}{2} \mathrm{tr}\log\delta^2 S+{\cal O}(\hbar^2), \label{loopexp} \end{equation} where \begin{equation} \delta^2 S(x,y)=\frac{\delta^2 S_k[\phi+\phi'_0]}{\delta\phi'(x) \delta\phi'(y)}\, , \end{equation} and the saddle point, $\phi'_0$, is defined by the extremum condition \begin{equation} \frac{\delta S_k[\phi+\phi'_0]}{\delta\phi'}=0, \end{equation} in which the infrared background field, $\phi(x)$, is held fixed. It can be proved that the saddle point is trivial, $\phi'_0=0$, as long as the matrix $\delta^2S(x,y)$ is invertible and the IR background field is homogeneous, $\phi(x)=\Phi$. Now, each successive loop integral in the $n$-loop contributions which are not explicitly written in Eq.~(\ref{loopexp}) brings a suppression factor \begin{equation} \frac{k^d-k^{\prime d}}{k^{\prime d}}={\cal O}\left(\frac{k-k'}{k'}\right) \end{equation} due to the integration volume in the momentum space. Thus $\delta k/k=(k-k')/k$ appears as a new small parameter which suppresses the higher loop contributions in the blocking relation and the ``exact'' functional differential equation obtained in the limit $\delta k\to 0$ includes the one-loop contribution only. But we should bear in mind that the loop expansion had to be used at the initial stage of the derivation so the resulting ``exact'' equation might be unreliable in the strong coupling situation. All we know is that the loop corrections to the evolution equation obtained in the one loop level are vanishing. We will use the gradient expansion for the action, \begin{equation} S[\phi]=\sum_{n=0}^\infty \int \mathrm{d}^d x\, U_n(\phi(x),\partial^{2n}), \end{equation} where $U_n$ is an homogeneous function of order $2n$ in the derivative. In the leading order of this expansion, the so-called local potential approximation, we have \begin{equation} S[\phi]=\int\mathrm{d}^dx \left[\frac{Z(\phi)}{2}(\partial_\mu\phi)^2+U(\phi)\right], \label{localpotapprox} \end{equation} and furthermore the simplification $Z(\phi)=1$ will be used to derive a simple differential equation for the potential $U$. This local potential will be then the only function characterizing the action. If we use now a homogeneous infrared background field, $\phi(x)=\Phi$, we obtain from Eqs.~(\ref{loopexp}) and~(\ref{localpotapprox}) an equation for the local potential $U_k(\phi=\Phi)$: \begin{equation} U_{k-\delta k}(\Phi)=U_k(\Phi)+\frac{1}{2}\mathrm{tr} \log[\square+U_k^{\prime\prime}(\Phi)]+{\cal O}(\delta k^2), \label{WHeq} \end{equation} where we have introduced the notation \begin{equation} U_k^{\prime\prime}(\Phi)=\frac{\partial^2 U_k(\Phi)}{\partial\Phi^2}, \quad \square=-\partial_\mu\partial^\mu \end{equation} and the trace is taken in the subspace of the eliminated modes. We can explicitly write the trace in momentum space and get \begin{equation} U_{k-\delta k}(\Phi)=U_k(\Phi)+\frac{1}{2}\int \frac{\mathrm{d}^d p}{(2\pi)^d}\log[p^2+U_k^{\prime\prime}(\Phi)]+{\cal O}(\delta k^2), \end{equation} where the integration extends over the shell $k-\delta k<p<k$. In the limit $\delta k\to 0$ one then finds the differential equation \begin{equation} k\frac{\partial}{\partial k}U_k(\Phi)=-\frac{\Omega_d k^d}{2(2\pi)^d} \log[k^2+U_k^{\prime\prime}(\Phi)], \label{WHeqcont} \end{equation} where $\Omega_d$ denotes the $d$-dimensional solid angle \begin{equation}\label{solidc} \Omega_d=\frac{2\pi^{d/2}}{\Gamma\left(\frac{d}{2}\right)}. \end{equation} The Wegner-Houghton equation~(\ref{WHeqcont}) represents the one-loop resummed mixing of the coupling constants of the potential $U_k=\sum_n (g_n/n!)\Phi^n$. In fact, an expansion of the logarithm in the second derivative of the potential gives \begin{equation} k\frac{\partial}{\partial k}U_k(\Phi)= -\frac{\Omega_d k^d}{2(2\pi)^d}\sum_{n=1}^\infty\frac{1}{n} \left(\frac{-U_k^{\prime\prime}(\Phi)}{k^2+U_k^{\prime\prime}(\Phi)} \right)^n, \end{equation} up to a field independent constant. This is the usual one loop resummation of the effective potential~\cite{coleman73} except that the loop momentum is now restricted to the subspace of the modes to be eliminated. Actually, the fact that the r.h.s. includes the running potential $U_k(\Phi)$ rather than the bare one, $U_\Lambda(\Phi)$, indicates that the contributions of the successive eliminations of the degrees of freedom are piled up during the integration of the differential equation and the solution of the renormalization group equation resums the perturbation series. The solution of the differential equation interpolates between the bare and the effective potential as $k$ is lowered from the original cutoff $\Lambda$ to zero. Finally, let us note that the derivation of Eq.~(\ref{WHeqcont}) shows that the restoring force for the fluctuations into the equilibrium is proportional to the argument of the logarithm function. Thus a nontrivial saddle point should be used when \begin{equation} k^2+U_k^{\prime\prime}(\Phi)\le0. \label{broken} \end{equation} \section{Evolution of the coupling constants} Our effective action is \begin{equation} S_k=\int \mathrm{d}^d x \left[\frac{1}{2}(\partial_\mu\phi(x))^2+ U_k(\phi(x))\right], \end{equation} and the initial condition for the evolution equation is given at $k=\Lambda$. The potential $U_k(\Phi)$ is assumed to be polynomial so it is expanded as \begin{equation} U_k(\Phi)=\sum_{n=0}^N \frac{1}{n!}g_n(\Phi_0)(\Phi-\Phi_0)^n. \label{expansion} \end{equation} We study the model in the symmetric phase, where the saddle point is trivial, $\Phi_0=0$. The polynomial structure of the potential is consistent because we avoid the singularity of Eq.~(\ref{WHeqcont}) at $k_\mathrm{cr}^2=-U_{k_\mathrm{cr}}^{\prime\prime}(\Phi)$ (recall Eq.~(\ref{broken})) which occurs in a region around $\Phi=0$ in the symmetry broken phase~\cite{polonyi97}. Taking the $n$-th derivative of Eq.~(\ref{expansion}) at $\Phi=0$ we obtain the coupling constant $g_n(k)$, and define the corresponding beta function by \begin{eqnarray} g_n(k)&=&\frac{\partial^n}{\partial\Phi^n}U_k(\Phi)_{|\Phi=0},\nonumber\\ \beta_n&=&k\frac{\mathrm{d}}{\mathrm{d} k}g_n(k)= \frac{\partial^n}{\partial\Phi^n}k\frac{\partial}{\partial k}U_k(\Phi). \end{eqnarray} By taking the successive derivatives of Eq.~(\ref{WHeqcont}), we obtain \begin{equation} \beta_n=-\frac{\Omega_d k^d}{2(2\pi)^d}\mathcal{P}_n(G_2,\ldots,G_{n+2}), \label{betafunct} \end{equation} where \begin{equation} G_n=\frac{g_n}{k^2+g_2} \label{bigg} \end{equation} and \begin{equation} \mathcal{P}_n=\frac{\partial^n}{\partial\Phi^n} \log[k^2+U_k^{\prime\prime}(\Phi)] \end{equation} is a polynom of order $n$ in the variables $G_j$, $j=2,\ldots,n+2$, \begin{eqnarray}\label{polinomia} \mathcal{P}_1&=&G_3, \nonumber \\ \mathcal{P}_2&=&G_4-G_3^2, \nonumber \\ \mathcal{P}_3&=&G_5-3G_3G_4+2G_3^3, \nonumber \\ \mathcal{P}_4&=&G_6-4G_5G_3-3G_4^2+12G_3^2G_4-6G_3^4,\\ \mathcal{P}_5&=&G_7-5G_6G_3-10G_5G_4+20G_5G_3^2+30G_4^2G_3-60G_4G_3^3+24G_3^5, \nonumber \\ \mathcal{P}_6&=&G_8-6G_7G_3-15G_6G_4-10G_5^2+30G_6G_3^2+ 120G_5G_4G_3+30G_4^3\nonumber\\ &&-120G_5G_3^3-270G_4^2G_3^2+360G_4G_3^4-120G_3^6.\nonumber \end{eqnarray} The coupling constants defined through Eq.~(\ref{expansion}) are dimensional parameters (the field variable $\phi$ has dimension $(d-2)/2$). However, the corresponding dimensionless parameters have more physical sense. We obtain them in the following way: \begin{equation} \tilde g_n(k)=k^{-n(1-\frac{d}{2})-d} g_n(k)= (\Lambda\tilde{k})^{-n(1-\frac{d}{2})-d} g_n(k), \label{dimensionless} \end{equation} where now $\tilde{k}$ runs from 1 to 0. Their beta functions are \begin{equation} \tilde\beta_n=-\left[n\left(1-\frac{d}{2}\right)+d\right]\tilde g_n +k^{-n(1-\frac{d}{2})-d}\beta_n, \label{bettil} \end{equation} where the first and the second term stands for the tree-level and the loop corrections, respectively. One can see that the super-renormalizable coupling constants follow asymptotically free scaling law at the tree level. \section{UV scaling laws and their extensions} One can distinguish an ultraviolet and an infrared scaling regime, for $k^2\gg |m^2(k)|$ and for $k^2\ll |m^2(k)|$, respectively. In the UV regime the scale dependence comes dominantly from the $k^2$ term of the propagator, see the denominator of Eq.~(\ref{bigg}); the $k$-dependence is generated by the phase factor $k^d$ in the IR regime where $k^2$ could be neglected in the inverse propagator. We will begin at the UV scale with the usual $\phi^4$ potential ($g_2\equiv m^2$) \begin{equation} V_\Lambda(\phi)=\frac{1}{2} m^2\phi^2+\frac{1}{4!}g_4\phi^4+\frac{1}{6!}g_6\phi^6, \end{equation} and see how the different couplings are generated when we move towards the IR regime. One ignores the $g_2$ term in the denominator of Eq.~(\ref{bigg}) in the asymptotic UV regime and finds \begin{eqnarray}\label{betafo} \frac{\mathrm{d} g_2}{\mathrm{d} k}&=&-\frac{\Omega_d}{2(2\pi)^d}\,k^{d-3}g_4,\nonumber\\ \frac{\mathrm{d} g_4}{\mathrm{d} k}&=&\frac{\Omega_d}{2(2\pi)^d}\,k^{d-3}\left( \frac{3}{k^2}g_4^2-g_6\right), \label{evolutions} \\ \frac{\mathrm{d} g_6}{\mathrm{d} k}&=&-\frac{\Omega_d}{2(2\pi)^d}\,3g_4k^{d-5}\left( \frac{10}{k^2}g_4^2-5g_6\right),\nonumber \end{eqnarray} where in the last equation we omitted the contribution of $g_8$. Consider the usual strategy in which the coupling constants $g_n$ are neglected for $n>4$ and the resulting equation is easy to integrate, \begin{equation} \frac{1}{g_4(k)}=\frac{1}{g_4(\Lambda)}+\frac{3\Omega_d \left(1-\left(\frac{k}{\Lambda}\right)^{4-d}\right)}{(4-d)2(2\pi)^dk^{4-d}}. \label{g4UVsc} \end{equation} This expression agrees with the result of the minimal subtraction (MS), a scheme which proved to be specially convenient in the ultraviolet scaling regime. It is based on the analytical continuation of the loop integrals in the ultraviolet domain so the resulting beta functions are mass independent, i.e. the terms ${\cal O}(g_2/k^2)$ are neglected. When extrapolating to the infrared regime we find erroneously the mass independent result $g_4\sim k^{4-d}$ ($g_4\sim \log k$ in $d=4$), $g_4$ tends to zero as $k\to 0$. This can be understood by inspecting (\ref{betafo}) where we find large positive values in the infrared for $g_4\not=0$ (this conclusion remains valid for finite $g_6$, as well). When the mass term is retained the beta function assumes the correct behavior and becomes ${\cal O}(k^d)$ in the infrared. Note that the term with $g_6$ acts in the opposite manner than $g_4$, c.f. the different signs in the right hand side of (\ref{betafo}), so that it can change the evolution considerably. It is instructive to look into the evolution of the dimensionless coupling constant, \begin{equation} \frac{1}{\tilde g_4(k)}=\frac{\left(\frac{k}{\Lambda}\right)^{4-d}} {\tilde g_4(\Lambda)} +\frac{3\Omega_d \left(1-\left(\frac{k}{\Lambda}\right)^{4-d}\right)}{(4-d)2(2\pi)^d}. \end{equation} For $4-d>0$ the one-loop $\omega=4-d$ universal critical exponent is reached for $k$ values sufficiently below the cutoff where $k/\Lambda\approx0$. The latter condition is needed to get rid of the non-universal cutoff effects. The scaling changes qualitatively as $d\to4$ because the non-universal $k\approx\Lambda$ behavior is spread over the whole $k$ range due to the smallness of $4-d$. This is what happens in the expansion \begin{equation} 1-\left(\frac{k}{\Lambda}\right)^{4-d}\to(d-4)\ln\frac{k}{\Lambda}, \end{equation} employed in the dimensional regularization scheme. This generalizes to any dimension: the marginal coupling constant follows the scaling law extended from the non-universal cutoff regime. The evolution for $g_2$ is of the form \begin{equation} \frac{\mathrm{d} g_2}{\mathrm{d} k}=-\frac{\Omega_d}{2(2\pi)^d}\,g_4 k^{d-3}. \label{gtwoas} \end{equation} It predicts \begin{equation} \frac{\mathrm{d} g_2}{\mathrm{d} k}\sim k,~~~g_2\sim k^2+\text{const} \end{equation} in lack of any dimensional constant. In the IR scaling regime we neglect the $k^2$ term in the denominator of Eq.~(\ref{bigg}) and using Eqs.~(\ref{betafunct}) and~(\ref{polinomia}), we get for $g_4$ (assuming again $g_6=0$), \begin{equation} \frac{\mathrm{d} g_4}{\mathrm{d} k}=\frac{3\Omega_d}{2(2\pi)^d}\,k^{d-1}\,\frac{g_4^2}{g_2^2}. \label{g4IRsc} \end{equation} This evolution is much slower comparing with Eq.~(\ref{g4UVsc}). In fact, for $d>1$ we have now a suppression factor $k^{d-1}$ which makes the coupling to stabilize at the attractive IR fixed point. In the same way, we obtain for $g_2$ \begin{equation} \frac{\mathrm{d} g_2}{\mathrm{d} k}=-\frac{\Omega_d}{2(2\pi)^d}\,k^{d-1}\,\frac{g_4}{g_2}, \label{g2IRsc} \end{equation} with a variation which is slower than that predicted by the UV scaling and a suppression factor for $k\to 0$. We see that the extraction of the scaling in the limit $k\to 0$ from the UV scaling laws, which is the commonly accepted practice in perturbation theory is incorrect when the MS scheme is used. One has to come back to the complete scaling laws in order to describe correctly the IR scaling. Four-dimensional asymptotically free gauge models present an infrared Landau pole in perturbation theory. But this behavior results from the extrapolation from the UV scaling laws. As we have just remarked, the IR limit of the UV regime is not correct in general, and the mass term can change considerably the actual behavior in the IR. Moreover, at the IR side of the UV regime, there are nonlinear effects that make important the contribution of the irrelevant (non-renormalizable) couplings (see Fig.~3 (b), commented in the next Section), therefore even the IR limit of the UV scaling can be influenced by these couplings. These ideas have been considered qualitatively in the previous paragraphs, after inspection of Eq.~(\ref{evolutions}). Let us now examine them more quantitatively. We take as an example of asymptotically free model the scalar theory in three dimensions. We know from the epsilon-expansion result~\cite{wilson72} that below four dimensions, the $\lambda\phi^4$ theory does not present an infrared Landau pole, but a fixed point located at \begin{equation} \tilde{\lambda}^*=\frac{16\pi^2}{3}\epsilon, \label{WFfixedpoint} \end{equation} at order $\epsilon=4-d$. This is the Wilson-Fisher fixed point. For finite $\epsilon$, for example in two and three dimensions, we can get this one-loop result from our beta functions of the dimensionless coupling constants obtained from Eqs.~(\ref{bettil}) and~(\ref{betafunct}) in the asymptotic UV regime (that is, ignoring $g_2$), which for the $\tilde{g}_4$ coupling give \begin{equation} \tilde{\beta}_4=-(4-d)\tilde g_4+ 3\tilde g_4^2\frac{\Omega_d}{2(2\pi)^d}, \end{equation} giving the IR fixed point \begin{equation} \tilde g_4^*=\frac{(4-d)2(2\pi)^d}{3\Omega_d}. \end{equation} Restricting ourselves to the $d=3$ case, we find \begin{equation} \tilde g_4^*=\frac{4\pi^2}{3}\simeq 13.15947. \label{g4FP} \end{equation} However, in three dimensions $g_6$ is a marginal coupling, and it can be generated by the RG flow, modifying the position of the fixed point~(\ref{g4FP}). If we include $\tilde g_6$ in our analysis, we get the beta functions \begin{eqnarray} \tilde{\beta}_4&=&-\tilde g_4+\frac{3}{4\pi^2}\tilde g_4^2- \frac{1}{4\pi^2}\tilde g_6 \, ,\nonumber \\ \tilde{\beta}_6&=&\frac{-3}{4\pi^2}\tilde g_4\,(10 \tilde g_4^2-5\tilde g_6). \end{eqnarray} It is immediate to see that the zeros of these beta functions are at the point \begin{eqnarray} \tilde g_4^*&=&4\pi^2\simeq 39.4784, \nonumber \\ \tilde g_6^*&=&32\pi^4\simeq 3117.091 . \label{fixpe} \end{eqnarray} But now let us consider the inclusion of a non-renormalizable coupling, $g_8$. The beta functions for the $(g_4,g_6,g_8)$ model are \begin{eqnarray} \tilde{\beta}_4&=&-\tilde g_4+\frac{3}{4\pi^2}\tilde g_4^2- \frac{1}{4\pi^2}\tilde g_6 \, ,\nonumber \\ \tilde{\beta}_6&=&\frac{-3}{4\pi^2}\tilde g_4\,(10 \tilde g_4^2-5\tilde g_6) -\frac{1}{4\pi^2}\tilde g_8 \, , \label{betag8} \\ \tilde{\beta}_8&=&\tilde g_8-\frac{7}{4\pi^2}\, (-90\tilde g_4^4+60\tilde g_4^2\tilde g_6-5\tilde g_6^2-4\tilde g_4\tilde g_8). \nonumber \end{eqnarray} From them, one obtains the fixed point \begin{eqnarray}\label{fixpk} \tilde g_4^*=\frac{4\pi^2}{210}\left(195 + \sqrt{29625}\right) \simeq 69.01, \nonumber \\ \tilde g_6^*=-4\pi^2 \tilde g_4^* + 3\tilde g_4^{*2}\simeq 11563, \label{fixed_point_g8} \\ \tilde g_8^*=15\tilde g_4^{*3}-60\pi\tilde g_4^{*2}\simeq 2.11\cdot 10^6. \nonumber \end{eqnarray} These values are also obtained in the numerical integration of Eqs.~(\ref{betag8}), independently of the initial values for the different couplings (if they are different from zero, which corresponds to the Gaussian fixed point). To assess the importance of this result, the difference between the physics around the fixed points (\ref{fixpe}) and (\ref{fixpk}), recall that the modification of the irrelevant operator set at the cutoff influences the overall scale of the model. Thus one has to consider dimensionless quantities in comparing the two coupling constant regions. The most obvious candidate, the dimensionless ratio between the mass and the four point vertex, $g_2/g^2_4$, is trivially vanishing in our approximation. But $g_6$ is dimensionless and its variation at the fixed points indicates that no adjustment of the overall scale could bring the physics of these two fixed points together. As we have expected, the non-renormalizable coupling $g_8$ modifies the position of the fixed point without changing the blocking procedure, turning to a situation of strong coupling dynamics in the IR. This is because the linearity which one assumes to ignore the irrelevant coupling constants is no longer valid in the strong coupling regime. As $g_4$ and $g_6$ approach their large IR fixed point the linearization fails and new scaling laws are found which in turn generate new relevant operators \cite{polonyi97}, overlapping with $\phi^8$. Thus the strong coupling dynamics may induce a new (and artificial) IR scaling regime even if the UV scaling laws are extrapolated down to low energies. Nothing unusual happens for infinitesimal $\epsilon$ when one stays in the vicinity of the Gaussian fixed point. In that case, linearity applies all along the RG flow from the Gaussian fixed point to the Wilson-Fisher fixed point given by Eq.~(\ref{WFfixedpoint}). However, as we have seen, its location is changed for finite $\epsilon$ by nonlinear effects produced in the flow from one fixed point to the other. In the vicinity of this infrared fixed point (which, we stress again, is artificial in the sense that it neglects the influence of the mass term) we of course have again the classification of relevant and irrelevant terms, which, in the case of three and two dimensions, is different from the one obtained from power counting. For example, this IR fixed point has in two dimensions just the mass and the fourth order coupling as relevant parameters, while higher order couplings are irrelevant. It is well known that the poles of the fixed point action at complex values of the field variable make the Taylor expansion in the field unreliable \cite{trpot}. We do not see any reason to reject a blocked action only because the potential is diverging beyond a given field strength. This kind of internal space singularity might only indicate a maximal particle density in the system. We should stay only sufficiently far from this limiting value of the field variable when the evolution equation is truncated. We interpret the difference of the two fixed points as an indication of the breakdown of the simple universality which is based on the linearized flow equation around the UV fixed point. So far we considered the extrapolation of the UV scaling laws to the IR regime. Does the conclusion concerning the importance of the non-renormalizable coupling constant $g_8$ remains valid when the true evolution equation, with $g_2\not=0$, is considered? The mass slows down the evolution but this may happen ``too late'' and the strong coupling effects can be found on the true renormalization group trajectory for small enough renormalized mass, close enough to the critical point. When the mass is large then crossover freezes the evolution of the coupling constants ``earlier'' and the linearization remains valid. To demonstrate this case recall that the IR limit of the UV regime means a fixed point for the dimensionless couplings. For a super-renormalizable coupling constant such as $g_4$, which has positive dimension, this would mean that the dimensional coupling goes to zero when $k \to 0$. However, we know that for a relevant coupling, the dimensionless quantity diverges when $k\to 0$, so that the dimensional coupling will take a finite value at the IR. This reasoning can be explicitly checked in Fig.~1, in which we consider the $d=3$ scalar theory with just one coupling, $g_4$. The white points follow the evolution of the UV regime and its extrapolation to $k=0$. We observe that indeed the dimensionless coupling reaches the fixed point given by Eq.~(\ref{g4FP}), while the dimensional coupling goes to zero. However, if one considers the complete beta function, i.e. retaining $g_2$ (black points), the behavior is the same in the UV regime, but then the true trajectory separates from the IR limit of this regime and enters into the actual IR regime, which implies a divergent dimensionless coupling at $k=0$, and a certain finite value of the dimensional coupling, as explained above. One can however see numerically that this finite value is stable and almost does not change when one introduces more and more non-renormalizable couplings in the RG evolution. This is what one expects when the crossover captures the coupling constants and slows down their evolution in the regime of linearizability where the non-renormalizable couplings are unimportant. In the same way, it might well be that the IR Landau pole observed in four-dimensional gauge theories is just an artifact of a wrong IR limit or the truncation of the renormalized action. First, the nature of the singularity can change when one adds non-renormalizable couplings, and then the IR Landau pole would be the reflection of the insufficient functional form of the blocked action, and second, the true IR trajectory can be quite different from the IR limit of the UV scaling. \section{Asymptotic freedom and the perturbation expansion} We examine in this Section the scaling laws in dimensions $d=2,3$ and 4 from the point of view of the applicability of the perturbation expansion. As we have seen in the previous Section, in the RG evolution of our model there are two asymptotic scaling regimes, $k\to\infty$ and $k\to0$. The latter one is trivial as mentioned above, because the beta functions \eq{betafunct} are suppressed by the factor $k^d$ and the evolution of the dimensional coupling constants slows down as $k\to0$. The asymptotic UV scaling is however more involved. The super-renormalizable (relevant) and the renormalizable (marginal) coupling constants, $g_n$ with $n<2d/(d-2)$ and $n=2d/(d-2)$ according to the power counting, respectively, follow their autonomous evolution, the universal renormalized trajectory\footnote{Ignoring the triviality in $d=4$ where the tree level marginal coupling constant $g_4$ is actually irrelevant due to the radiative corrections.}. The non-renormalizable (irrelevant) coupling constants ``forget quickly'' their initial value and take values which are generated by the universal flow. This general trend is demonstrated by the renormalization group flow shown for $d=4$ and 3 in Figs.~2 and~3, respectively. In those Figures the dimensionless coupling constants are displayed as the functions of the cutoff which is measured in the units of the initial cutoff value, $k\to k\Lambda$. In the case of theories with non-Gaussian asymptotically free couplings (the case of our model for $d<4$), an excessive growth of these couplings in the UV regime may produce a non-perturbative situation in the infrared. We want to study this by comparing the values of the couplings at $k=0$. To do so, we will adopt the convention that a model with positive renormalized mass square is non-perturbative in the vertex $g_n\phi^n$ if the radiative correction ${\cal O}(g_n)$ to the self energy is stronger than the mass term, i.e. \begin{equation} \frac{\tilde g_n(k)}{((n-2)/2)! \, 2^{(n-2)/2}\tilde m^2(k)}\gg1 \label{pert_cond} \end{equation} where $\tilde m^2(k)=\tilde g_2(k)$. In case this inequality were satisfied in the infrared ($k=0$), this would mean a non-perturbative situation and the invalidity of renormalized perturbation expansion. We remark that we are asking here about the validity of the perturbative condition at the true, $k=0$, infrared fixed point, where the RG flow ends. We are not considering for example the situation at the Wilson-Fisher IR fixed point in $d<4$, which, as was explained in Section 4, is the IR limit of the UV scaling behavior (a fixed point which describes the behavior of the system at the end of the UV regime), and not the real IR fixed point if we ask for the behavior of the system at energy scales much lower than the mass $m(k)$ (for example, in dimension two, we will see that at the crossover, or the end of the UV regime, the high order couplings start to take large values, while they are irrelevant couplings for the Wilson-Fisher IR fixed point; this is because the true RG flow separates from the extrapolation of the UV flow, as was explicitely shown in Fig. 1 in the case of $d=3$). We now turn to a detailed analysis of the situation at dimensions four, three and two. Since the neglected higher order vertices may influence the evolution while we lower the cutoff, we have to address the problem of the system of coupled equations numerically. The evolution of a non-renormalizable coupling constant, $\tilde g_6$ in four dimensions, is shown in Fig.~2. The irrelevance is expressed by the independence of $\tilde g_6(k)$ on the initial value $\tilde g_6(1)$ for $k\ll1$\footnote{The different initial value for the non-renormalizable coupling constants may induce a different overall scale factor. This effect is very weak in our case due to the smallness of the renormalizable coupling constants.}. The value given in the leading order of the perturbation expansion,\footnote{Which is not applicable for the strongly coupled case~\eq{fixpe} or~\eq{fixpk}.} ${\cal O}(\tilde g_4^3)$, is reached at $k\approx0.3$ for not too large values of $\tilde g_6(1)$. Since the model is in the weak coupling regime the evolution is rather slow after arriving at this universal value if the cutoff is high enough to provide a long scaling regime. In our case the scale window $0.3<k<1$ was insufficient and the plateau is reduced into a peak at $k\approx0.3$ before the crossover. But the bringing of $\tilde g_6(1)$ close to the universal value creates a plateau even with this limited range of the scales as it can be seen for $\tilde g_6(1)=9\cdot10^{-8}$. Such a scale independence is the ultimate goal of the improved action program~\cite{impr}. As the initial value increases beyond the plateau level the coupling constant decreases in a monotonous manner as the cutoff is lowered. In order to check the critical exponent coming from the linearized blocking we need a $\tilde g_6(k)$ larger than the universal value since the latter originates in the nonlinear level, ${\cal O}(\tilde g_4^3)$. The evolution follows the linear relation $\tilde g_6\approx k^2$ for $k<1$ indicated by the dashed line in the last plot of Fig.~2. The infrared scaling regime is $0<k<0.1$ where $\tilde g_6(k)$ tends to zero with $k$ along the universal trajectory. The theory remains perturbative in $d=4$ since it has no relevant non-Gaussian coupling constant. The three-dimensional renormalization group flow is depicted in Fig.~3. The asymptotic infrared scaling laws are rather simple, the super-renormal\-iz\-able coupling constants diverge, the renormalizable one $n=6$ converges and the non-renormalizable ones tend to zero in the infrared, $k\to0$. The ultraviolet scaling law, above $k\approx0.1$ indicates the weak, radiative correction generated relevance of $\tilde g_6$ for the given initial conditions, $\tilde g_n(1)$. The insensitivity on the initial condition $\tilde g_8(1)$ for $k<0.3$ seen in the last plot supports the irrelevance of $\tilde g_8$. In fact, the evolution of $\tilde g_8(k)$ follows the linearized scaling law as long as its value is far form the universal ${\cal O}(\tilde g_4^4)$ value. It is worthwhile noting that the non-renormalizable coupling constants $\tilde g_n(k)$ always develop a peak of sign $(-1)^{1+n/2}$ around the crossover, $k\approx0.1$. The appearance of the peak can be understood as the result of the increase of $|\tilde g_n(k)|$ from zero as the cutoff is lowered in the ultraviolet scaling regime and the decrease in the infrared side of the crossover. Non-perturbative phenomena may arise at the low energy edge of the UV scaling regime due to the increase of the asymptotically free coupling constants, $g_4$ and $g_6$ if the scaling regime is long enough and the initial value of the coupling constants $\tilde g_4(1)$ and $\tilde g_6(1)$ are large enough. There is however no problem in finding a perturbative, asymptotically free theory in the infrared. The parameter with the highest energy dimension in the Lagrangian is $g_2=m^2$ for $d>2$. Thus $\tilde g_2$ is the largest among the dimensionless coupling constants according to Eq.~\eq{bettil}, and it dominates the action and renders the theory perturbative in the IR limit. The comparison of this conclusion with Eq.~\eq{fixed_point_g8} reveals the necessity of treating the IR scaling laws properly in establishing the validity of the renormalized perturbation expansion. As $d$ approaches 2 more and more coupling constants become super-renormalizable. The fastest increasing dimensionless non-Gaussian coupling constant during the decrease of the cutoff is $\tilde g_4$. The critical exponent, the measure of the speed of the increase become degenerate for infinitely many coupling constants when $d=2$. The specialty of the lower critical dimension is the existence of infinitely many super-renormalizable coupling constants, $\tilde g_n$, with equal critical dimension. This degeneracy of the dimensions evolves the non-Gaussian pieces of the action with the same rate as the mass term on the tree level and the theories are not obviously perturbative any more. In other words, it remains for the radiative corrections, the last term in the right hand side of Eq.~\eq{bettil} to determine if the theory runs into weak- or strong-coupling regime at the infrared. The renormalization group equations were integrated out numerically in two dimensions with the initial conditions $(g_2^i=-0.001, g_4^i=0.01, g_n^i=0, n>4)$ to find the evolution of the coupling constants. The result, depicted in Fig.~4, shows a marked increase at the low energy end of the UV scaling regime. This increase originates from the asymptotically free evolution. The relevant behavior of a coupling constant is defined on the linearized level of the blocking, i.e. in the leading order of the perturbation expansion. This one-loop result can be obtained by replacing the running coupling constants in the beta functions by their initial values at the cutoff. The local potential obtained in the one-loop approximation is \begin{eqnarray} U_k(\phi)&=&V_\Lambda(\phi)+\frac{1}{2}\int_{k<p<\Lambda} \frac{\mathrm{d}^2 p}{(2\pi)^2} \log[p^2+V_\Lambda^{\prime\prime}(\phi)]= \nonumber \\ &&\frac{1}{8\pi}\left\{[\Lambda^2+V_\Lambda^{\prime\prime}(\phi)] \log[\Lambda^2+V_\Lambda^{\prime\prime}(\phi)]- \right. \nonumber \\ &&\left.[k^2+V_\Lambda^{\prime\prime}(\phi)] \log[k^2+V_\Lambda^{\prime\prime}(\phi)]-\Lambda^2+k^2\right\}, \label{oneloop} \end{eqnarray} with \begin{equation} V_\Lambda^{\prime\prime}(\phi)=m^2(\Lambda)+\frac{g_4(\Lambda)}{2}\phi^2. \end{equation} The comparison of the numerical solution with the one-loop evolution is shown in Fig.~5. The one-loop formula~(\ref{oneloop}) cannot be extended down to $k=0$, because there will be a value of $k$ such that $k^2+m^2(\Lambda)=0$, since $m^2(\Lambda)$ is negative. But we only want to compare the results in the perturbative regime. So we have stopped the evolution in Fig.~5 at $k\sim 0.3$. One can also see an increase in the one-loop solution at small $k$ which accumulates and drives the system non-perturbative at lower values of $k$. The numerical results of Fig.~4 show that the initial conditions $(g_2^i=-0.001, g_4^i=0.01, g_n^i=0, n>4)$ correspond to a non-perturbative system. Can we find initial conditions which yield perturbative dynamics? In order to answer this question the left hand side of the inequality \eq{pert_cond} is plotted against the initial value for $g_2$ on Fig.~6, for the different couplings up to $N=20$, at a value of $g_4^i=0.001$. The result does not change qualitatively for different values of bare $g_4$. It supports the general trend of having the systems more perturbative when the Gaussian part of the action is increased. The higher order coupling constants tend to grow faster but it seems that $g_n$ can be brought into the perturbative regime for sufficiently large initial mass square. At $g_2^i=0.01$, for example, all the couplings are perturbative, and this perturbative character is more pronounced for the high couplings. However, the separation between the values of the l.h.s. ratio of Eq.~\eq{pert_cond} at $g_2^i=0.01$ is smaller as we go to higher couplings, and one can ask whether it has got a limiting value or the trend can be reversed for a sufficiently high order coupling. Indeed, Fig.~7 reveals that this happens for the $n=24$ coupling ($N=26$), which suggests that one will have a non-perturbative situation also at this value of bare $g_2$ going to a sufficiently high order coupling constant. The situation is the same, even stronger, below two dimensions: the existence of an infinite number of relevant couplings makes that one cannot assure perturbativity by looking to a finite number of couplings, no matter what the initial conditions are. We take this and similar other failures in finding a perturbative theory observed at different initial conditions as a strong numerical indication of the non-perturbative nature of {\em any} two and lower dimensional scalar field theory with polynomial interaction. \section{Lattice Wegner-Houghton equation} \label{lattice} The previous Sections dealt with the renormalization group flow at finite scales. We address now a different, asymptotic problem, the manner the sensitivity on the initial values of the irrelevant coupling constants is suppressed during the renormalization. This question is usually rendered trivial by the universality argument. But there are two reasons to suspect that such a reasoning which is based on the linearization of the blocking relation might be oversimplified; both are related to an infinite set of operators. The reason motivating a more careful check of the universality, mentioned in the Introduction, is that the models at or below the lower critical dimension contain infinitely many relevant operators. It is not obvious whether the sum over the interaction vertices is always convergent enough to make the linearization of the blocking relation a reliable approximation. Another potential problem shows up if one changes infinitely many irrelevant terms in the action by choosing another regulator. Let us compare the momentum space cutoff in the continuum with the lattice regularization. The propagator is a monotonic function of the momentum in the continuum. This is not the case on the lattice. In fact, the fermion doubling problem on the lattice~\cite{nogo} results from the periodicity of the propagator in the first Brioulline zone, the appearance of $2^d-1$ new maxima in the propagators in the UV, non-universal regime. The existence of a maximum of the propagator in the UV regime contradicts an assumption of the studies of the continuum models, namely that the propagator decreases monotonically as $p\to\infty$, and renders the perturbation expansion non-universal for lattice fermionic models~\cite{reisz}. There is no species doubling for bosons but their propagator remains periodic on the lattice and we find $2^d-1$ lattice extrema in the UV regime. The existence of these extrema is an evidence of the slowing down in the decrease of the propagator as the momentum approaches the boundary of the first Brioullin zone. This in turn indicates the weaker suppression of the high energy modes compared to the continuum regularization. Does this mean that the UV scaling laws are different in the continuum than on the lattice? We shall find an affirmative answer to this question but this result does not contradict the universality. Some words of caution are in order at this point. One would object the interpretation of the modification of the cutoff as the introduction of new irrelevant coupling constants by recalling that the theory ceases to be renormalizable in the presence of the non-renormalizable (irrelevant) couplings. The resolution of the apparent paradox is based on the difference between the ways the renormalization group is used in Statistical and High Energy Physics. We are interested in the dynamics close to the cutoff in Statistical Physics and this is respected by the employment of the blocking which keeps the {\em complete} dynamics unchanged below the actual cutoff. The price of this precision is the appearance of the infinitely many irrelevant coupling constants in the action. We seek the dynamics at finite, fixed scales in High Energy Physics. Since the cutoff is sent to the infinity this boils down the problem of keeping the physics cutoff independent {\em far from the cutoff} only. The obvious gain of such an ease of the conditions is the freedom from the adjustment of the non-renormalizable parameters. Thus one can remove the cutoff when the non-renormalizable parameters are present in the action without any problem\footnote{Ignoring again the possibility of the triviality, the appearance of an UV Landau pole.} as long as the renormalization conditions are imposed far from the cutoff. The lattice regularization of the scalar model can be described by using the momentum space as the introduction of the non-renormalizable higher order derivative terms, \begin{eqnarray} (\partial_\mu\phi)^2&\longrightarrow& \frac{4}{a^2}(\sin\frac{a\partial_\mu}{2i}\phi)^2\\ &=&\left(\sum\limits_{\ell=0}^\infty \frac{1}{(2\ell+1)!}\left(\frac{a}{2}\right)^{2\ell}\partial^{2\ell+1} \phi\right)^2.\nonumber \end{eqnarray} The cutoff dependence of the non-renormalizable coupling constants follows a tree-level relation arising from the Taylor expansion of the sine function. This is sufficient to establish convergent physics at finite scales when $a\to0$~\cite{reisz}, a claim to be verified in this Section numerically by means of the implementation of the Wegner-Houghton scheme on the lattice. But this convergence can not rule out a modification of the scaling laws in the asymptotical UV regime. In fact, we shall find a new scaling regime between the region where the usual universal UV scaling is observed and the UV fixed point. The only effect the different adjustments of the non-renormalized coupling constants may leave on the finite scale physics can be comprised in an overall scale factor. It is rather straightforward to repeat the steps leading to Eq.~\eq{WHeqcont} on the lattice. It is shown in the Appendix that the only change required is the modification of the ``solid angle'' factor, $\Omega(k)$: the lattice evolution equation \eq{WHeqlatt} is obtained from Eq.~\eq{WHeqcont} by replacing Eq.~\eq{solidc} by Eq.~(\ref{omega3}). One recovers the continuum solid angle for $d=2$, $\Omega_2=2\pi$, in Eq.~(\ref{omega3}) as $k\to0$, thus the WH equations agree in the IR limit. In fact, one sees numerically that the behavior of the evolution of the different coupling constants in the lattice RG is qualitatively the same as in the continuum case. But the question we are interested in is the relation between the regularizations in the UV, where the coupling constants are introduced, when the physics is the same at finite scales. It is shown in the Appendix that there is a natural relation between the cutoffs, $\Lambda^2=8/a^2$, which matches the finite scale physics. We shall follow the renormalization group flow in terms of the coupling constants whose dimension is removed by the initial value of the cutoff, \begin{equation} g_n \longrightarrow g_n/\Lambda^2, \label{rescaling} \end{equation} in order to avoid the singularities at $k=0$. Let us consider $\lambda\phi^4$ lattice theory which can be studied either numerically or analytically and whose properties can be matched to those of the continuum theory by the adjustment of $g_2$ and $g_4$. But the situation is more involved in two dimensions. The reason is again that there are infinitely many renormalizable coupling constants and one cannot match the finite scale physics by adjusting $g_2$ and $g_4$. This is demonstrated in Fig.~8 where the lattice model with the initial conditions $m^2(8)=g_2(8)=-0.001$, $g_4(8)=0.01$, and $g_n(8)=0,~n>4$ (where we have already made the rescaling Eq.~(\ref{rescaling})) was evolved in the infrared direction. As the system reached $k=k_\mathrm{end}$ the continuum WH equation was used to increase the cutoff. The result is a ``perfect matching'' of the models in the UV which gives the same low energy physics in the IR. As we can see in Fig.~8, the lattice $\lambda\phi^4$ model in two dimensions is {\em not} the continuum $\lambda\phi^4$ theory. It contains contributions of infinitely many renormalizable other coupling constants. Of course, numerically we had to truncate the equations at a certain coupling (here, at ${\cal O}(\Phi^{22})$) but we checked that these ``truncation effect'' hardly influences the values of the low order coupling constants. We have taken for the parameter $k^2_\mathrm{end}$ the value $k^2=0.3$ in Fig.~8, while the crossover is at $k^2_\mathrm{cr}\sim 0.01$. We had to use $k^2_\mathrm{end}>k^2_\mathrm{cr}$, because the high order couplings have very large values at the crossover which requires very fine discretization in the numerical resolution of our differential equations to ensure that the way back to the UV is done accurately. However, $k^2_\mathrm{end}$ should also be sufficiently small that the flow be universal there, in other words to make sure that the irrelevant lattice contributions are suppressed for $k^2<k^2_\mathrm{cr}$. The choice $k^2_\mathrm{end}>0$ introduces an uncertainty in the matching. To assess it we repeated the ``go-return'' evolution described above and checked the discretization errors for $k^2_\mathrm{end}=0.8,~0.5$ and $0.3$. After then we took the appropriate bare parameters at the UV end points in both regularizations and followed the evolutions down to $k=0$. The relative difference, \begin{equation} \Delta g_n\equiv \frac{|g_n^\mathrm{CONT}(0)-g_n^\mathrm{LATT}(0)|}{|g_n^\mathrm{LATT}(0)|} \label{relative} \end{equation} is shown in Table~1 for the different coupling constants at $k=0$. The smallness of the deviation assures that the IR behavior has practically been obtained with $k^2_\mathrm{end}=0.3$, and one can trust the conclusions, the approach of the flow to a universal curve, extracted from Fig.~8. The first two plots in Fig.~8 show that the mass and the quartic coupling constant run parallel in the SUV region of the lattice regularization (see Appendix) and in the continuum. There is no convergence between the two regularizations in this unusual scaling regime, anticipated above. The approach to the universal curve starts for $k^2<4$, below the SUV regime only. The fact that the renormalization group flow converges to the universal one in the $2^{-d}$-th part of the Brioullin zone only sets an unexpected high lower limit on the lattice size when the continuum limit is sought in numerical simulations. The higher order vertices seem to converge to the universal curve from the very beginning but the difference between the two regularizations is surprisingly large. The universal trajectory of the $\phi^4$ model is reached later by the higher order vertices. This effect appears to be a counterpart of the non-perturbative features seen in Figs.~6,~7 and introduces a large uncertainty in identifying two-dimensional models in different regularizations. \section{Conclusions} The renormalization group flow of scalar models with polynomial interaction is considered in the first part of this paper by solving the Wegner-Houghton equation numerically in the local potential approximation for $d=2,~3$ and 4. The numerical results showed in this paper suggest that the length of the UV scaling regime which is needed to generate non-perturbative dynamics in the infrared shrinks to zero as the number of the asymptotically free coupling constants tends to infinity. In other words, the $\Lambda$-parameter tends to the cutoff as the lower critical dimension is approached, $d\to2$. Such a behavior limits considerably the values of the coupling constants for a perturbative system in dimension 3 and renders {\em all} two- and lower dimensional field theories with polynomial couplings non-perturbative. This makes the understanding of the noncritical low dimensional condensed matter systems more involved. The one-dimensional models belong to first quantized quantum mechanics and our result is a manifestation of the failure of the convergence of the perturbation expansion for an anharmonic oscillator. Such a conclusion does not invalidate the well known results for two-dimensional systems, such as the applicability of the Bethe ansatz, bosonization and the availability of certain exact information for models with conformal invariance. Instead, it makes the asymptotic state structure and the relation between the the dressed particles and the states created by the application of the field operator from the vacuum more involved. We found an interesting analogy between the infrared Landau pole of the confining four-dimensional Yang-Mills theories and the low dimensional scalar models which opens the possibility of an unexpected, nontrivial structure in the asymptotic states in the low dimensional scalar models. Viewed with interest in particle physics our conclusion suggests that one can avoid the IR Landau pole by following the evolution of the non-renormalizable operators. {\em How to find the non-renormalizable operators whose presence stabilises the theories at low energies?} It is well known that massive Lagrangians generate trivial infrared scaling laws, i.e. the Gaussian mass term is the only relevant operator in the infrared scaling regime. This is because the fluctuations are exponentially suppressed beyond the correlation length so the evolution of the coupling constants slows down at the infrared side of the crossover. The theories with dimensional transmutation, i.e. dynamically generated scale parameter or infrared instability only can support non-perturbative dynamics in the IR scaling regime. Thus the operators sought should be relevant in the IR regime, their growth being fed by IR or collinear divergences. There are few known cases only where the low modes are controlled by non-renormalizable operators. These include the four fermion contact term in solids inducing the BCS transition~\cite{shankar}, the higher order derivative terms in the action which generate inhomogeneous vacuua~\cite{afvac}, the common element being the onset of a Bose-Einstein condensation~\cite{polonyi97}. In the second part of the paper the infinitesimal renormalization group scheme is generalized for lattice regularization. The matching of the continuum and the lattice regularizations is carried out numerically and the approach of the universal renormalization group flow is demonstrated for the two-dimensional $\phi^4$ lattice model. This result suggests that the naive argument for the universality, which is based on the linearization of the blocking relations remains valid in the presence of infinitely many relevant operators. Other potential troublemakers, the infinitely many higher order derivatives contained in the lattice kinetic energy do generate a new, ``super UV'' scaling regime but universality is restored at the IR end of the usual UV scaling regime. Another use of the lattice regulated version of the Wegner-Haughton equation is the estimate of the finite size effects in a non-perturbative manner. This provides a useful check of the thermodynamic limit of the numerical results obtained in general on small lattices. \section*{Appendix} The details of the derivation of the Wegner-Houghton equation in the lattice regularization are given in this Appendix. Let us consider the scalar field theory regularized on a lattice of lattice spacing $a=1$. We want to derive a WH equation similar to Eq.~(\ref{WHeqcont}). We integrate over spherical shells in the momentum space for the continuum regularization, because the propagator has spherical symmetry. This is no longer the case on the lattice, where we have \begin{equation} \square=\sum_{\mu=1}^d\hat p_\mu^2, \quad \hat p_\mu=2\sin\frac{p_\mu}{2}. \end{equation} Let us see the surfaces of equal value of the lattice propagator in two dimensions by performing the following change of variables: \begin{equation} \renewcommand{\arraystretch}{1.5} \left. \begin{array}{c} (p_x,p_y)\longrightarrow (p,\theta) \\ 4\sin^2\frac{p_x}{2}+4\sin^2\frac{p_y}{2}=p^2 \\ \tan\theta=\frac{\sin (p_y/2)}{\sin (p_x/2)} \end{array} \right\} \label{chvar} \end{equation} We can see in Fig.~9 the form of the curves of constant propagator for several values of $p^2$. $p$ can be identified as the ``momentum scale'' that runs from the cutoff at $p^2=\Lambda^2=8$ to the IR $p^2=0$. It is also clear in that figure that the value $p^2=4$ separates two regimes, still in the ultraviolet region, that we could call super-UV (SUV), for $8>p^2>4$, and normal-UV regimes. For $p^2\sim 0$, the lines are spheres, and our change of variables~(\ref{chvar}) reduces to the usual relation between cartesian and polar coordinates. The absolute value of the Jacobian of the transformation~(\ref{chvar}) is found to be \begin{equation} J=\frac{J_\mathrm{p}}{\sqrt{\left(1-\frac{p^2}{4}\cos^2\theta\right) \left(1-\frac{p^2}{4}\sin^2\theta\right)}}, \end{equation} where $J_\mathrm{p}$ is the usual Jacobian for the polar change of variables, $J_\mathrm{p}=p$. The transformation~(\ref{chvar}) can be easily generalized to three and four dimensions; however, we can only treat analytically the integral that appears in the derivation of the WH equation in the $d=2$ case. To derive the equivalent of the Wegner-Houghton equation~(\ref{WHeqcont}) in the bidimensional lattice regularization we will start from Eq.~(\ref{WHeq}), and calculate the trace by integrating in momentum space over a shell $k-\delta k<p<k$, where $p$ is the parameter we have introduced in Eq.~(\ref{chvar}), \begin{eqnarray} \frac{1}{2}\mathrm{tr}\log[\square+U_k^{\prime\prime}]&=& \frac{1}{2}\int\frac{\mathrm{d}^2 p}{(2\pi)^2}\log[4\sin^2\frac{p_x}{2}+ 4\sin^2\frac{p_y}{2}+U_k^{\prime\prime}]= \nonumber \\ &&\frac{1}{2(2\pi)^2}\int\mathrm{d}\theta\int_{k-\delta k}^k \mathrm{d} p\,\, J\, \log[p^2+U_k^{\prime\prime}]\approx \nonumber \\ &&\frac{\delta k}{2(2\pi)^2} k \log[k^2+U_k^{\prime\prime}]\,\Omega(k), \end{eqnarray} with \begin{equation} \Omega(k)=\int\mathrm{d}\theta\, \frac{8}{\sqrt{64-16 k^2+k^4\sin^2 2\theta}}. \label{omegageneral} \end{equation} We have to distinguish two different regimes in making the integration: \medskip (i) $\underline{k^2<4.}\quad$ In this region the range of values for $\theta$ is $(0,2\pi)$. \begin{equation} \Omega(k)=\int_0^{2\pi}\mathrm{d}\theta\, \frac{8}{\sqrt{64-16 k^2+k^4\sin^2 2\theta}}= \sum_{i=1}^4\Omega_i(k), \end{equation} where we have split the interval $(0,2\pi)$ into four intervals $(0,\pi/2)$, $(\pi/2,\pi)$, etc. Let us consider $\Omega_1(k)$. With the change of variable $x=\tan\theta$ and the notation \begin{equation} \bar k^2=4-k^2, \end{equation} this integral can be brought into the form \begin{equation} \Omega_1(\bar k)=\frac{1}{\bar k}\int_0^\infty \frac{2\,\mathrm{d} x}{\sqrt{(x^2+b^2)(x^2+b^{-2})}}, \label{integral1} \end{equation} where $b^2=4/\bar k^2$. The integral in Eq.~(\ref{integral1}) is related to the elliptic integral of the first kind~\cite{tablas80} $F(\phi,t)$, \begin{equation} \Omega_1(k)=F\left[\frac{\pi}{2},\frac{k}{4}\sqrt{8-k^2}\right];\quad F(\varphi,t)=\int_0^\varphi\frac{\mathrm{d}\alpha}{\sqrt{1-t^2\sin^2\alpha}}. \end{equation} The result is the same for the other integrals $\Omega_i$, $i=2,3,4$. But $F(\pi/2,t)$ is a complete elliptic integral, which can be expressed in terms of the hypergeometric function~\cite{tablas80,tablas65} \begin{equation} F(\alpha,\beta;\gamma;z)=\frac{\Gamma(\gamma)}{\Gamma(\alpha)\Gamma(\beta)} \sum_{n=0}^\infty\frac{\Gamma(\alpha+n)\Gamma(\beta+n)}{\Gamma(\gamma+n)} \frac{z^n}{n!} \label{hypfunct} \end{equation} as \begin{equation} \Omega(k)=2\pi F\left(\frac{1}{2},\frac{1}{2};1;\frac{(8-k^2)k^2}{16}\right). \label{omega1} \end{equation} \medskip (ii) $\underline{4<k^2<8.}\quad$ This is the SUV region. We split again the integral into four integrations in the corresponding quadrants. By using the same change of variables as above, we have to calculate \begin{equation} \Omega_i(\bar k)=\int\frac{4\,\mathrm{d} x}{\sqrt{x^2(\bar k^4+16)-4\bar k^2(1+x^4)}}, \end{equation} where \begin{equation} \bar k^2=k^2-4. \end{equation} Special care is needed at the limits of integration (recall Fig.~9). It can be seen that the four integrals can be put together in the form \begin{eqnarray} \Omega(\bar k)&=&4\int_{\bar k/2}^{2/\bar k} \frac{4 \,\mathrm{d} x}{\sqrt{x^2(\bar k^4+16)-4\bar k^2(1+x^4)}}=\nonumber \\ &&\frac{8}{\bar k}\int_{\bar k/2}^{2/\bar k} \frac{\mathrm{d} x}{\sqrt{(x-\bar k/2)(2/\bar k-x)(x+\bar k/2)(x+2/\bar k)}}. \end{eqnarray} This integral is related again~\cite{tablas80} to an elliptic integral and an hypergeometric function: \begin{equation} \Omega(k)=\frac{32}{k^2}F\left[\frac{\pi}{2},\frac{8}{k^2}-1\right]= \frac{16\pi}{k^2}F\left(\frac{1}{2},\frac{1}{2};1,\frac{(8-k^2)^2}{k^4}\right). \label{omega2} \end{equation} \medskip We would like to have a common expression for $\Omega(k)$ for both cases (i) and (ii). From Eqs.~(\ref{omega1}) and~(\ref{omega2}), we find in fact that the expressions differ in a factor 2 for $k^2=4$! The reason is that actually our integral is divergent at this point. From Eq.~(\ref{omegageneral}) we see that the divergent integral is \begin{equation} \Omega(k^2=4)=\int_0^{2\pi}\mathrm{d}\theta\, \frac{2}{|\sin 2\theta|} \end{equation} (in fact, the hypergeometric function~(\ref{hypfunct}) converges in general only in the unit circle $|z|<1$~\cite{tablas80}). We will see, however, that this divergence is integrable during the RG evolution from $k^2=8$ to $k^2=0$, and therefore it has no physical significance. In order to have a consistent, single expression for the cases $8<k^2<4$ and $4<k^2<0$, we will make use of the following property of the hypergeometric functions~\cite{tablas80}: \begin{equation} F\left(\frac{1}{2},\frac{1}{2};1,z^2\right)= \frac{1}{1+z}F\left(\frac{1}{2},\frac{1}{2};1,\frac{4z}{(1+z^2)^2}\right), \quad 0\leq z<1. \label{prop} \end{equation} Let us consider the expression~(\ref{omega1}) which is valid for $k^2<4$. Using the property~(\ref{prop}) one finds, \begin{equation} \Omega(k)=2\pi F\left(\frac{1}{2},\frac{1}{2};1, \frac{4z}{(1+z^2)^2}\right)= 2\pi(1+z)F\left(\frac{1}{2},\frac{1}{2};1;z^2\right), \label{useofprop} \end{equation} where we have set \begin{equation} \frac{4z}{(1+z^2)^2}=\frac{(8-k^2)k^2}{16}. \end{equation} This equation has two solutions for $z$ as a function of $k$: \begin{equation} z=\quad \frac{k^2}{8-k^2}\quad; \quad \frac{8-k^2}{k^2}, \end{equation} but only the first one is admissible for Eq.~(\ref{useofprop}), because it gives $z<1$ for $k^2<4$, while the second solution gives a value greater than 1 in this region. Now, if we define $\tilde{k}=8-k^2$, then we have $z=(8-\tilde{k}^2)/\tilde{k}^2$, and Eq.~(\ref{useofprop}) becomes \begin{equation} \Omega(\tilde{k})=2\pi\frac{8}{\tilde{k}^2} F\left(\frac{1}{2},\frac{1}{2};1;z^2\right). \end{equation} Comparing this last expression with the result~(\ref{omega2}) for the case $8<k^2<4$, we can finally write \begin{eqnarray} \Omega(k)&=&\frac{16\pi}{\bar k^2} F\left(\frac{1}{2},\frac{1}{2};1;z^2\right), \text{ with } z=\frac{8-\bar k^2}{\bar k^2} \nonumber \\ &\text{and}&\begin{cases} \bar k^2=k^2 & \text{if $k^2=8\ldots 4$},\\ \bar k^2=8-k^2 & \text{if $k^2=4\ldots 0$}. \end{cases} \label{omega3} \end{eqnarray} The hypergeometric function $F\left(\frac{1}{2},\frac{1}{2};1;z^2\right)$ can be computed directly from its definition~(\ref{hypfunct}). One can obtain a high precision in the evaluation of the series with a reasonable number of terms (say, around 50) when $z$ is not very close to 1, say, for $0.7>z>0$. For $1>z>0.7$ we have used the following alternative formula~\cite{tablas65} \begin{eqnarray} F(\alpha,\beta;\alpha+\beta;z)&=& \frac{\Gamma(\alpha+\beta)}{(\Gamma(\alpha)\Gamma(\beta))^2} \sum_{n=0}^\infty \frac{\Gamma(\alpha+n)\Gamma(\beta+n)}{(n!)^2} [2\psi(n+1)\nonumber\\ &&-\psi(\alpha+n)-\psi(\beta+n)-\log(1-z)](1-z)^n, \nonumber \\ &&(|\mathrm{arg}(1-z)|<\pi, |1-z|<1), \end{eqnarray} where~\cite{tablas65} \begin{equation} \psi(z)=\frac{\mathrm{d} \,\log\Gamma(z)}{\mathrm{d} z}, \end{equation} which gives a better convergence for the function $F\left(\frac{1}{2},\frac{1}{2};1;z^2\right)$ near $z=1$ because it is a series in the variable $(1-z^2)$. In conclusion, our generalization of the WH equation~(\ref{WHeqcont}) for a lattice regularization in two dimensions is \begin{equation} k\frac{\partial}{\partial k}U_k(\Phi)= -\frac{\Omega(k) k^2}{2(2\pi)^2}\log[k^2+U_k^{\prime\prime}(\Phi)], \label{WHeqlatt} \end{equation} where $k$ is the parameter $p$ of Eq.~(\ref{chvar}), and $\Omega(k)$ is given by Eq.~(\ref{omega3}). \section*{\protect Acknowledgements} We wish to thank V. Branchina, S.B. Liao, J. Alexandre, H. Mohrbach, E. Vicari and A. Pelissetto for useful discussions. Work partially supported by the Spanish MEC, Acci\'on Integrada hispano-francesa HF1997-0041, the French program, Actions Integr\'ees franco-espagnol, Picasso 98064. J.M.C. acknowledges support from the EU TMR program ERBFMRX-CT97-0122. He also thanks the Spanish MEC, the CAI European program and DGA (CONSI+D) for financial support.
\section{Neutrino Masses} These lectures discuss how the universe serves as a learning ground for massive neutrinos. Before doing so, let us briefly review some experimental measurements of neutrino masses. Upper bounds on neutrino masses from kinematic measurements in laboratories continue to improve.~\cite{mass} For the $\tau$-neutrino, $m_{\nu_\tau}< 18.2$ MeV from the decay channel $\tau \rightarrow 5\pi + \nu_\tau$. For the $\mu$-neutrino, $m_{\nu_\mu}< 170$ keV from two-body pion decay. For the electron neutrino, the quantity $m_{\nu_e}^2$ is measured in tritium beta decay by fitting the shape of the energy spectrum near the endpoint. Experiments thus far have yielded nonphysical negative values for $m_{\nu_e}^2$, indicating unexplained systematic effects in the measurements. A conservative upper bound is put at $m_{\nu_e}\approx 15$ eV. The spread in arrival times of neutrinos from supernova explosions provides an independent way to constrain the mass of the electron neutrino. Various limits have been reported for SN 1987A; a conservative estimate is $m_{\nu_e}< 23$ eV.~\cite{mass,mann} \section{Properties of Cosmic Background Neutrinos} \subsection{Temperature and Density} For a brief 1 second after the big bang, neutrinos enjoy being part of the thermal bath composed of photons, electrons, protons, neutrons, and the associated anti-particles (after the quark-hadron era). The weak interactions at this early time are rapid enough to keep these particles in thermal equilibrium at a single temperature $T$. After 1 second, when $T$ drops below about 1 MeV, however, the neutrino interaction rate becomes slower than the Hubble expansion, and neutrinos become effectively collisionless and freely-streaming particles whose trajectories are determined by the geodesic equations. This event is commonly referred to as ``neutrino decoupling.'' As the universe expands, the momenta and temperature of neutrinos are simply redshifted, and the neutrino temperature is given by the familiar formulas \begin{equation} T_\nu(a) = a^{-1}\,T_{\nu ,0}\,,\qquad T_{\nu ,0}=\left( {4\over 11}\right)^{1/3} T_{\gamma ,0} =1.947 K \,, \end{equation} where $a$ is the cosmic scale factor, the subscripts 0 denote the present-day values, and the cosmic background photon temperature is taken to be $T_{\gamma ,0}=2.728\,K$.~\cite{fix96} An important feature of the neutrino distribution after decoupling is that, although weak interactions are no longer rapid enough to keep neutrinos in thermal equilibrium with other particle species, neutrinos retain their equilibrium distribution as long as no other physical processes (e.g., gravitational clustering; see Sec.~3) are present to alter it. Therefore, to zeroth order in density and metric perturbations, the phase space distribution $f_0$ of the cosmic background neutrinos is of the simple Fermi-Dirac form \begin{equation} f_0(\epsilon)={g_s\over h_p^3} {1\over e^{\epsilon/ k_B\,T_{\nu,0}}+1}\,, \label{fermi} \end{equation} where $\epsilon=a(p^2+m_\nu^2)^{1/2}$ is the comoving energy, $T_{\nu,0}$ is the neutrino temperature given by Eq.~(1), $g_s$ is the number of spin degrees of freedom, and $h_p$ and $k_B$ are the Planck and the Boltzmann constants. The situation is further simplified if neutrino masses are $\ll 1$ MeV. Such neutrinos are highly relativistic at decoupling; their energy $\epsilon$, and hence the distribution function $f_0$, are independent of $m_\nu$ to a good approximation. One can easily show that, as long as $m_\nu\ll 1$ MeV, the number density of the cosmic background neutrinos is related to the neutrino temperature by \begin{equation} n_\nu(T_\nu) = {7g_s\over 8\pi^2}\zeta(3) \left( {k_B\,T_\nu\over \hbar\,c}\right)^3 \,, \end{equation} where $\zeta(3)\approx 1.202$ is the Riemann zeta function of order 3. This gives a present-day density of $\approx 113$ cm$^{-3}$ for every neutrino species independent of their masses. (For comparison, the present-day photon density is $\approx 412$ cm$^{-3}$.) It also follows that the contribution of these neutrinos to the present-day mass density parameter, $\Omega_\nu$, is related to their masses by the simple relation \begin{equation} \Omega_\nu\,h^2 = {\Sigma_i m_i \over 93\, {\rm eV}}\,, \end{equation} where the index $i$ runs over all light, stable neutrino species (e.g., $\nu_e, \nu_\mu$, and $\nu_\tau$), and the Hubble constant is $H_0=100\,h$ km s$^{-1}$ Mpc$^{-1}$. One then arrives at the important conclusion that in order for neutrinos not to close universe (i.e. $\Omega_\nu\le 1$), the sum of neutrino masses must not exceed $93\,h^2$ eV. This value is far below the current laboratory limits (see Sec.~1). Cowsik \& McClelland~\cite{cm72} were the first to use such cosmological arguments to place an upper bound on neutrino masses. (Unfortunately, these ``hot dark matter'' models in which the mass density is dominated by massive neutrinos have been found to produce excessive large voids surrounded by large coherent sheets and filaments that are not seen in the observable universe.~\cite{white83} Modifications to this model will be discussed below.) In the high mass regime, $m_\nu \gg 1$ MeV, there exists another window where the neutrino contribution to the mass density parameter $\Omega$ of the universe is subcritical. The argument is that neutrinos with $m_\nu \gg 1$ MeV become non-relativistic long before decoupling. Neutrino and anti-neutrino pairs cease to be created in abundance once the thermal temperature drops below $m_\nu$, and the neutrino density is suppressed by the Boltzmann factor $e^{-m_\nu/k_B\,T}$. This large reduction factor in the relic abundance allows neutrinos to have large masses without overclosing the universe. A more careful calculation~\cite{lee} shows that an $\Omega\le 1$ universe implies a {\it lower} limit of $\sim 2$ GeV if these heavy neutrinos are Dirac, and $\sim 6$ GeV if they are Majorana. Since this mass range is well above the current upper mass bounds from laboratory measurements, it is of interest only when one considers more exotic theories for neutrinos.~\cite{pgl} \subsection{Kinematics and Free Streaming} Let us now turn to the kinematics and the streaming properties of neutrinos. In general, neutrinos of mass $m _\nu$ become non-relativistic after a redshift of \begin{equation} z_{\rm rel} \approx {m_\nu c^2 \over 3k_B\,T_{\nu ,0}} = 2\times 10^3 \left( {m_\nu \over 1\, {\rm eV}} \right) \,. \end{equation} This redshift has important implications for structure formation because it dictates the time at which massive neutrinos begin to make a transition from being radiation to matter. Note that this transition occurs fairly early, before recombination if $m_\nu\go 1$ eV. The average momentum of the cosmic background neutrinos at temperature $T_\nu$ is given by \begin{equation} \left< p \right > = 3.15\,k_B\, T_\nu/ c\,. \end{equation} In the non-relativistic regime ($p=m_\nu\,v$), the average neutrino speed can be written as \begin{equation} \left< v \right> = 160\, {\rm km/s\ } \left( {1\, {\rm eV}\over m_\nu} \right) \left( {T_\nu\over 1.947} \right) \,. \label{vave} \end{equation} Since $T_\nu \propto a^{-1}\propto (1+z)$, massive neutrinos slow down as time goes on. It is important to keep in mind that neutrinos with a mass of several eV have slowed down to an average velocity below 100 km s$^{-1}$ today. We also note that at the redshift of matter-radiation equality, $z_{\rm eq} \sim 24000\,\Omega\,h^2$ (i.e. when the total energy density in radiation in the universe equals that in matter), light neutrinos with $1 < m_\nu < 10$ eV are zooming around with speeds close to $c$. Such large thermal speeds prevent massive neutrinos from clustering gravitationally during this epoch, and this is why light neutrinos are referred to as hot dark matter (HDM). In contrast, perturbations in cold dark matter (CDM), which by definition has negligible thermal velocities, can grow unimpeded after $z_{\rm eq}$. I will quantify the different clustering behavior of CDM and HDM further in Sec.~3 and 4. Since neutrinos cannot cluster appreciably via gravitational instabilities on scales below the free streaming distance, this introduces a characteristic length scale into the problem. This scale is given by the free-streaming wavenumber (in comoving coordinates) \begin{equation} k_{\rm fs}^2= {4\pi G\rho\, a^2\over \left< v \right>^2 }\,, \end{equation} which is analogous to the Jeans length for a self-gravitating system of density $\rho$. For $k<k_{\rm fs}$ (i.e. large wavelengths) , the density perturbation in the neutrinos is Jeans unstable and grows unimpeded in the matter-dominated era. For $k>k_{\rm fs}$, the density perturbation decays due to neutrino phase mixing. A phase-space interpretation of the free streaming property is that the phase mixing of collisionless particles damps the growth of density perturbations.~\cite{bond83} When the neutrinos are relativistic, $\left< v\right> \approx c\,$, and the free-streaming distance is approximately the particle horizon, which scales as $k_{\rm fs}(a) \propto a^{-1}$ (in the radiation-dominated era). After the neutrinos become non-relativistic the relations $\left< v\right> \propto a^{-1}m_\nu^{-1}$, $m_\nu\propto\Omega_\nu h^2$, and $\rho\propto a^{-3} h^2$ then imply~\cite{ma96} \begin{equation} k_{\rm fs}(a)\propto a^{1/2}\Omega_\nu h^3\,. \label{kfs2} \end{equation} As expected, the free-streaming distance ($\propto k_{\rm fs}^{-1}$) decreases with time as the neutrinos slow down. This also implies that neutrinos can cluster gravitationally on increasingly small length scales at later times.~\cite{ma96} Such behavior has been seen in cosmological numerical simulations and will be discussed in Sec.~4. \section{Linear Perturbations in Neutrinos and Other Particles} Thus far our discussion has focused on the properties of the smooth cosmic background neutrinos, and Eqs.~(1)-(7) were derived under this assumption. The universe today, however, is clearly far from being homogeneous on scales of $\sim 100$ Mpc and below. Baryons and dark matter in galaxies, clusters, and superclusters show a wide spectrum of overdensities above the cosmic mean. The current theoretical framework for the origin and evolution of these cosmic structures rests upon the assumption that certain primordial fluctuations (perhaps originated from quantum fluctuations of scalar fields during the inflationary era) imprint a perturbation spectrum on all matter and radiation. These fluctuations subsequently grow via gravitational instabilities to give rise to the wide range of observed structures. How are the cosmic relic neutrinos affected by all this? To understand the growth of density perturbations in neutrinos as well as other forms of matter and radiation, one would need to learn the linear cosmological perturbation theory of gravitational instability. A full description of this theory requires more time than is allocated for these lectures. I will only sketch the theory below with emphasis on the neutrino component. Interested readers should refer to the pioneering work of Lifshitz, later reviewed in Lifshitz \& Khalatnikov.~\cite{lifshitz} More modern treatments of various aspects of this theory can be found in the textbooks by Weinberg and Peebles,~\cite{text} in the reviews by Kodama \& Sasaki and Mukhanov, Feldman \& Brandenberger,~\cite{review} and in the Summer School lectures by Efstathiou, Bertschinger, and Bond.~\cite{school} A complete description of this theory for all relevant particles is given by Ma \& Bertschinger.~\cite{mb95} Here, I will only discuss the essence of the theory and highlight the physical meaning of the key results. \subsection{Neutrino Phase Space} Let us start with the neutrinos. The full phase space distribution function of neutrinos can be written as \begin{equation} f(\vec{x},\vec{p},t)=f_0(p)+f_1(\vec{x},\vec{p},t)\,, \end{equation} where $f_1$ denotes perturbations to the Fermi-Dirac distribution $f_0$ given by Eq.~(\ref{fermi}). Unlike the unperturbed term $f_0$ that depends only on $p$, $f_1$ can have complicated dependence on time as well as positions $\vec{x}$ and the conjugate momenta $\vec{p}$. The equations for neutrino temperature, number density etc. discussed in Sec.~2 were obtained assuming $f=f_0$. A non-vanishing $f_1$ would lead to perturbations in these quantities. For example, the perturbed neutrino energy density is related to $f_1$ by \begin{equation} \delta\rho(\vec{x},t)= a^{-4}\int d^3 p\,\epsilon\,f_1(\vec{x},\vec{p},t) \,. \end{equation} Other quantities such as perturbations in the pressure and shear can also be related to $f_1$. It is in general difficult to compute and sample $f_1$ directly because at a given time, it depends on six variables. A Monte Carlo technique, or a ``general-relativistic $N$-body'' technique, has been developed to evolve $f_1$ from redshift $z\sim 10^9$ shortly after neutrino decoupling until $z\sim 10$ when nonlinear effects become non-negligible.~\cite{mb94a} In this calculation, an ensemble of neutrino simulation particles is initially assigned velocities drawn from the Fermi-Dirac distribution, which is an excellent approximation at $z\sim 10^9$. The trajectory of each neutrino simulation particle is then followed by integrating the geodesic equations in the {\it perturbed} background spacetime. The metric perturbation gives rise to a nonzero $f_1$, and the particle positions and velocities at a later time $t$ represent a realization of $f_1(\vec{x},\vec{p},t)$. Results from this calculation have revealed that at $z\sim 15$, positive correlations have developed in the rms neutrino velocities and the overdensity, which would be absent if the phase-space distribution were purely Fermi-Dirac (i.e. $f=f_0$). The more spatially clustered neutrinos are found to move faster, possibly resulting from an increase in the kinetic energy during gravitational infalls. \subsection{Evolution of Perturbed Density Fields} The phase-space description above is applicable to all particle species and is used in the full theory. The full, general-relativistic version of the linear perturbation theory is described by a set of coupled and linearized Einstein, Boltzmann, and fluid equations. The variables include the metric perturbations to the homogeneous and isotropic Friedmann-Robertson-Walker spacetime, and the phase-space perturbations in all relevant particle species (e.g., photons, baryons, cold dark matter, massless and massive neutrinos). The Einstein equations describe how the time evolution of the metric perturbations is affected by the perturbations in the density, pressure, shear, and higher-order moments of matter and radiation. The Boltzmann and fluid equations, on the other hand, describe the time evolution of the radiation and matter distribution in the perturbed spacetime. Together, this theory describes the growth of metric and density perturbations throughout the early history of the universe, and it serves as the foundation for all calculations of the linear power spectra for matter and temperature variations imprinted on the cosmic microwave background. \begin{figure} \epsfxsize=5.truein \epsfbox{del.ps} \caption{Time evolution of the perturbed energy density field, $\delta=\rho/\bar{\rho}-1$, for five matter and radiation components in a flat C+HDM cosmological model. (See text for model parameters.) The results are from integration of the coupled Einstein and Boltzmann equations. Since the equations are linearized, each $k$-mode evolves independently. Two modes are shown here for illustration. In each panel, the five curves represent $\delta$ for the cold dark matter (solid), baryons (dash-dotted), photons (long-dashed), massless neutrinos (dotted), and massive neutrinos (short-dashed), respectively.} \label{fig:del} \end{figure} Figure~\ref{fig:del} illustrates a small subset of results that can be obtained from numerical integration of these linearized equations. It shows the time evolution of the perturbed energy density field, $\delta=\delta\rho/\bar{\rho}=\rho/\bar{\rho}-1$, for the five relevant particle species in a cold+hot dark matter (C+HDM) model. This model assumes an Einstein-de Sitter universe containing a mixture of CDM, HDM (i.e. massive neutrinos), baryons, photons, and massless neutrinos. The density parameters in the first three components in this model are $\Omega_c=0.75$, $\Omega_\nu=0.2$, and $\Omega_{\rm b}=0.05$, and the Hubble parameter is taken to be $H_0=50$ km s$^{-1}$ Mpc$^{-1}$, or $h=0.5$. From Eq.~(4), these parameters correspond to a neutrino mass of 4.7 eV. (For definiteness, this calculation has assumed that only one type of neutrinos, presumably $\nu_\tau$, has a non-negligible mass.) Two wavenumbers, $k=0.01$ Mpc$^{-1}$ (top) and 1 Mpc$^{-1}$ (bottom), are shown in Figure~\ref{fig:del} to demonstrate the intricate dependence of $\delta$ on length and time scales. The overall normalization of $\delta$ is set arbitrarily. We observe several salient features in Figure~\ref{fig:del}. First, the amplitudes of $\delta$ for all particles grow monotonically until a critical time, after which different particle species exhibit very different behavior. This critical time is the ``horizon crossing'' time, and it occurs when the horizon has grown large enough to encompass the wavelength of a given mode of perturbation. A mode of perturbation is therefore not in causal contact until horizon crossing. Naturally, this occurs earlier for smaller wavelengths (i.e. larger $k$). In Figure~\ref{fig:del}, one can indeed see that the $k=1$ Mpc$^{-1}$ mode enters the horizon at $a\sim 10^{-5}$ while the $k=0.01$ Mpc$^{-1}$ mode enters the horizon a little after $a= 10^{-3}$. A point to keep in mind is that the behavior of $\delta$ before horizon crossing is strongly dependent on the choice of gauge. The results shown in Figure~\ref{fig:del} are computed in the so-called synchronous gauge, which is a popular choice due to historical precedent.~\cite{lifshitz} See Mukhanov, Feldman \& Brandenberger~\cite{review} and Ma \& Bertschinger~\cite{mb95} for discussion of a more convenient gauge (the conformal Newtonian gauge). The second feature in Figure~\ref{fig:del} to note is that after horizon crossing, the photons (long-dashed) and baryons (dot-dashed) exhibit rapid, coupled oscillations in the $k=1$ Mpc$^{-1}$ mode but only the photons oscillate in the $k=0.01$ Mpc$^{-1}$ mode. This occurs because the former enters the horizon before recombination at $a_{\rm rec} \sim 10^{-3}$, and the photons and baryons are coupled by Thomson scattering and oscillate acoustically. The coupling is not perfect. The friction of the photons dragging against the baryons leads to Silk damping,~\cite{silk68} which is prominent in the bottom panel of Figure~\ref{fig:del} at $a\sim 10^{-3.5}$. After recombination, the baryons decouple from the photons and fall quickly into the potential wells formed earlier by the CDM. This results in the rapid growth of the dot-dashed curve in the bottom panel of Figures~\ref{fig:del}. The mode with $k=0.01$ Mpc$^{-1}$, on the other hand, enters the horizon when the universe is neutral. The baryons therefore grow like the CDM and do not oscillate. The critical length scale demarcating these two regimes is the horizon size at recombination $a_{\rm rec}$: $k_{\rm rec} \sim 0.03$ Mpc$^{-1}$. The third feature to note in Figure~\ref{fig:del} is the rate of growth of the CDM component (solid curve) after horizon crossing. Close inspection shows that the CDM in the bottom panel grows more slowly at $10^{-5}\lo a \lo 10^{-4}$ than later on, whereas in the upper panel, the CDM simply grows with a power law after horizon crossing. This is because the shorter wavelength mode ($k=1$ Mpc$^{-1}$) enters the horizon when the energy density of the universe is dominated by radiation, and fluctuations in matter (e.g. CDM) cannot grow appreciably during this era. The critical scale separating continual and suppressed growth is the horizon size at the time of radiation-matter equality $a_{\rm eq} \sim 4\times 10^{-5} (\Omega h^2)^{-1}$: $k_{\rm eq} \sim 0.1$ Mpc$^{-1}$ for the parameters of this model. The fourth feature to note in Figure~\ref{fig:del} is the behavior of the massive neutrinos (short-dashed). As discussed in Sec.~2, neutrinos of masses within the cosmologically interesting range ($\sim 1$ to 10 eV) are highly non-relativistic today but were relativistic at earlier times. This property is in fact evident in the top panel of Figure~\ref{fig:del}: Careful inspection shows that at $a\approx 10^{-4}$, the density field $\delta$ in massive neutrinos is indeed making a gradual transition from the upper line for the radiation fields to the lower line for the matter fields. (More precisely, the primordial perturbations are assumed to be ``isentropic'' here, which leads to a perturbation amplitude that is a factor of $4/3$ higher for radiation than matter.) The subsequent evolution of $\delta$ in massive neutrinos for this mode ($k=0.01$ Mpc$^{-1}$) is very similar to that of CDM. This is because it enters the horizon when the thermal velocities of the neutrinos have decreased substantially; the free-streaming effect is therefore unimportant. For the $k=1$ Mpc$^{-1}$ mode, on the other hand, the free streaming effect is evident and the growth of $\delta$ in massive neutrinos is suppressed until the characteristic free-streaming scale $k_{\rm fs}(a)$ given by Eq.~(9) grows to $\sim k$. Afterwards, the short-dashed curve for massive neutrinos is seen to grow again and catch up to the CDM. Since $k_{\rm fs} \propto a^{1/2}$, the larger $k$ modes suffer more free-streaming damping and $\delta$ for massive neutrinos can not grow until later times. The damping in the massive neutrino component also affects the growth of the CDM, slowing it down more for models with larger $\Omega_\nu$ compared to the pure CDM model. \subsection{Growth Rate and Power Spectrum} I have used Figure~\ref{fig:del} computed for two particular $k$-modes in a particular cosmological model to illustrate the physical meaning of many key features in the evolution of the density field for matter and radiation throughout the cosmic history. I will now discuss general descriptions that can be conveniently used to characterize the fluctuation amplitudes over a range of length scales for a variety of cosmological models. For example, it is extremely useful to know the dependence of $\delta$ and its time derivative on $k$ at a given time for a wide range of models. The most basic quantity to use is the linear power spectrum, $P(k,t)$, and the growth rate of the density field, $f\equiv d\log\delta/d\log a$. ({\em Note: I am following the convention of using $f$ to denote the growth rate; it should not be confused with the phase-space distribution function of Sec.~2 and 3.1.}) The power spectrum quantifies the two-point statistics of $\delta$, and for a Gaussian field, $P(k)$ represents its rms fluctuations and completely specifies its statistical properties. The power spectrum is therefore of fundamental importance in cosmology. \begin{figure} \epsfxsize=3.7truein \epsfbox[0 244 482 653]{f.ps} \caption{Growth rate of the CDM density field, $f\equiv d\log\delta/d\log a$, in four flat C+HDM models at $a=1$ (solid) and 0.1 (dashed). The four models assume different neutrino masses: $m_\nu=1.2$, 2.3, 4.6, and 6.9 eV (from top down), corresponding to $\Omega_\nu=0.05$, 0.1, 0.2, and 0.3. At small $k$, the CDM density field in these models grows with the same rate ($\delta\propto a$) as in the standard CDM model. At large $k$, the growth rate is suppressed because a fraction of the energy density in C+HDM models is in the hot neutrinos that exhibit less gravitational clustering. The suppression at large $k$ becomes less severe at later times because the velocities of the hot neutrinos decrease with time.} \label{fig:f} \end{figure} Comparing the growth rate and power spectrum for models with and without massive neutrinos is an effective way to illustrate the effects of hot dark matter. In the standard CDM model with $\Omega=1$ and $h=0.5$ (neutrino mass is assumed to be zero), the CDM density field grows as the expansion factor $a$ on all scales; therefore $f=1$. As discussed in Sec.~2.2, massive neutrinos introduce an additional length scale, the free-streaming distance, below which fluctuations are washed out and the growth rate is retarded. The growth rate is therefore more intricate in models with massive neutrinos and is generally a function of the wavenumber $k$, neutrino density parameter $\Omega_\nu$, and time. Figure~\ref{fig:f} illustrates such dependence in four different C+HDM models that assume a mixture of CDM and HDM. It shows that the growth is suppressed at large $k$, and models with a larger fraction of energy density in HDM suffer more. It also shows that the suppression becomes less severe at later times. One can gain some understanding of the behavior shown in Figure~\ref{fig:f} by exploring two asymptotic regions that can be solved analytically for C+HDM models with $\Omega_c+\Omega_\nu=1$: (1) Since HDM behaves like CDM above the free-streaming distance, $f\rightarrow 1$ as $k\rightarrow 0$; (2) In the opposite limit of large $k$, the HDM density field $\delta_h$ is severely dampened compared to the CDM density field $\delta_c$ because of the neutrino effects. For $\delta_h \ll \delta_c$, the time evolution of the CDM density field is governed by the linearized fluid equation \begin{equation} \ddot{\delta_c}+{\dot{a}\over a}\dot{\delta}_c=1.5 H^2 a^2\Omega_c \delta_c\,, \end{equation} where the dots denote differentiation with respect to the conformal time $\tau$. Since $Ha=2/\tau$ in the matter-dominated era, the growing solution in this regime is easily shown to be~\cite{bond80} \begin{equation} f_\infty\equiv f(k\rightarrow\infty) ={1\over 4}\sqrt{1+24\Omega_c}-{1\over 4} ={5\over 4}\sqrt{1-{24\over 25}\Omega_\nu}-{1\over 4}\,. \label{finfy} \end{equation} It is interesting to note that \begin{equation} f_\infty \approx \Omega_c^{0.6} \end{equation} is an excellent approximation to the equation above, especially for the cosmologically viable range of $\Omega_\nu\lo 0.3$. Using these analytic solutions in the asymptotic regimes and the scaling dependence of the free streaming wavenumber $k_{\rm fs}$ in Eq.~(\ref{kfs2}), one can construct a simple approximation for $f$ for a wide range of model parameters. It is found that the growth rate $f$ is well approximated by~\cite{ma96} \begin{equation} f \equiv {d\log \delta\over d\log a} = {1+ \Omega_c^{0.6}\, 0.00681\,x^{1.363} \over 1+0.00681\,x^{1.363}}\,,\qquad x\equiv {k\over \Gamma_\nu h} \,, \label{f} \end{equation} where $\Gamma_\nu$ is a shape parameter derived from Eq.~(\ref{kfs2}), \begin{equation} \Gamma_\nu=a^{1/2}\Omega_\nu h^2\,, \end{equation} $\Omega_c+\Omega_\nu=1$, and $k$ is in units of Mpc$^{-1}$. Note that Eq.~(\ref{f}) depends only on the variable $x$ that characterizes the neutrino free-streaming scale, and $\Omega_\nu$ (or $\Omega_c$) via $f_\infty$. The fractional error of the fit relative to the numerically computed values is smaller than 0.5\% for a wide range of parameters. The seemingly complicated multi-parameter dependence of Figure~\ref{fig:f} is succinctly incorporated in Eq.~(\ref{f}). \begin{figure} \epsfxsize=3.6truein \epsfbox[0 130 482 653]{p.ps} \caption{Power spectrum at $z=1.5$ for the perturbed matter density field for three flat C+HDM models with $\Omega_\nu=0.1$, 0.2, and 0.3. The solid curves show the linear $P(k)$ computed from the linear perturbation theory; accurate analytical fitting formulas are given by Eq.~(\ref{g}). The triangles show the nonlinear $P(k)$ computed from $N$-body simulations; accurate analytical approximations are given by Eq.~(\ref{master}).} \label{fig:pow} \end{figure} For the linear power spectrum, the slower time growth of $\delta$ at $k > k_{\rm fs}$ for the C+HDM models shown in Figure~\ref{fig:f} indicates a suppressed clustering amplitude on these scales. This effect is illustrated in Figure~\ref{fig:pow}, which shows the density-averaged power spectrum (i.e. $P=\{\Omega_c \sqrt{P_c}+\Omega_\nu \sqrt{P_\nu}\}^2$) for three flat C+HDM models. In general, $P_\nu \ll P_c$ on small length scales due to the neutrino thermal velocities, and the models with higher $\Omega_\nu$ clearly have less power at large $k$ in accordance with Figure~\ref{fig:f}. A good approximation for the relative $P(k)$ in a flat C+HDM model (with $\Omega_\nu\lo 0.3$) and a pure CDM model $(\Omega_\nu=0$) is given by~\cite{ma96} \begin{eqnarray} && {P(k,a,\Omega_\nu)\over P(k,a,\Omega_\nu=0)} = \left( { 1+b_1\,x^{b_4/2}+b_2\,x^{b_4} \over 1+b_3\,x_0^{b_4} } \right)^{\Omega_\nu^{1.05}}\,,\nonumber\\ && x={k\over \Gamma_\nu} \,, \quad x_0=x(a=1) \,, \label{g} \end{eqnarray} where the best-fit parameters are $b_1=0.004321, b_2=2.217\times 10^{-6}, b_3=11.63$, and $b_4=3.317$ for $k$ in units of Mpc$^{-1}$. Analytical approximations for the separate cold and hot spectra $P_c$ and $P_\nu$ can be found in the same reference. More complicated approximations for a wider range of parameter space have also been proposed.~\cite{eh98} \section{Nonlinear Gravitational Clustering of Neutrinos} As far as cosmological structure formation is concerned, the main difference between massive and massless neutrinos is that the former can participate in the processes of gravitational clustering and hence serves as a component of the dark matter in the universe. Massless neutrinos, on the other hand, affect cosmology only through their contribution to the radiation energy density. Any primordial perturbations in massless neutrinos are damped out after horizon crossing as a result of phase mixing (see dotted curves in Figure~\ref{fig:del}), and the only remnant of this component is the elusive $T_{\nu ,0}=1.947$ K background described in Sec.~2. \subsection{Spatial Distribution of Neutrinos} For massive neutrinos in models with a mixture of CDM and HDM, it is interesting to ask: do the neutrinos fall in the CDM potential wells and form a part of dark matter halos? One may naively think not because neutrinos are too hot. Our discussion thus far, however, indicates otherwise. We have seen in Eq.~(\ref{vave}) that neutrinos slow down with the expansion of the universe. Those with a mass of several eV are travelling with a speed much below the typical velocity dispersions $\sim 200$ km s$^{-1}$ of stars in galaxies. They can potentially be bound to galactic halos. In general, the extent to which massive neutrinos can cluster gravitationally depends on their mass and speed. \begin{figure} \epsfxsize=5.truein \epsfbox[80 144 592 718]{halo15.ps} \caption{Spatial distribution of cold dark matter (middle) vs. hot dark matter (bottom) particles in a simulated halo formed in a large $N-$body run for the flat $\Omega_\nu=0.2$ C+HDM model. The top panels shows the sum of cold and hot particles. Three redshifts are shown: $z=2$, 1, and 0 (from left to right), corresponding to cosmic times of 2.5, 4.6, and 13 billion years. All boxes have the same physical scale, $3.5\times 3.5\,h^{-1}$ Mpc. The final halo at $z=0$ is clearly a merger product of two dominant subhalos and several smaller satellites at higher redshifts. The spatial distribution of the massive neutrinos (4.7 eV) is visibly smoother than that of the CDM.} \label{fig:halo} \end{figure} To illustrate this point further, let us examine results from numerical simulations. Figure~\ref{fig:halo} shows the projected spatial distribution of cold particles (middle), hot particles (bottom), and the sum of the two (top) in a simulated dark matter halo in a flat $\Omega_\nu=0.2$ C+HDM model ($m_\nu=4.7$ eV). Three redshifts are shown (from left to right): $z=2, 1$ and 0 when the universe is 2.5, 4.6, and 13 billion years old, respectively. Each panel is $3.5\times 3.5\,h^{-1}$ Mpc in {\it physical} coordinates. The parent simulation is a large $N$-body run with 23 million simulation particles in a (100 Mpc)$^3$ comoving box.~\cite{mb94b} The dense halo shown at $z=0$ is clearly formed from mergers of two smaller halos and their satellites at higher redshifts, demonstrating the ``bottom-up'' hierarchical pattern of structure formation which is preserved in C+HDM models with $\Omega_c > \Omega_\nu$. Massive neutrinos are visibly clustered in the bottom panels, but their spatial distribution is smoother than that of the CDM. In the smaller CDM clumps at $z\go 1$, there are no discernible HDM halos at the same locations. The average thermal speed of an ensemble of 4.7 eV Fermi-Dirac neutrinos at $z\go 1$ is $\go 70$ km/s (cf. Eq.~\ref{vave}). As discussed in Sec.~3.1, when linear perturbations are taken into account, neutrinos in overdense regions have even higher velocities. It is therefore not surprising that the shallower potential wells of these small halos cannot trap a substantial number of HDM particles. The clustering of the 4.7 eV neutrinos shown in Figure~\ref{fig:halo} should be compared with the constraint derived by Tremaine \& Gunn,~\cite{tg79} which states that if cosmic neutrinos were to make up the bulk of galactic and cluster halos, they must be more massive than $\sim 10$ eV so that the Pauli exclusion principle is not violated. The C+HDM models considered here assume $\Omega_c > \Omega_\nu$ so as to preserve the successful hierarchical formation of structure in pure CDM models. Dark matter halos such as in Figure~\ref{fig:halo} have a substantial fraction of CDM, which in turn enhances the gravitational infalls of massive neutrinos. The clustering of neutrinos in C+HDM models is therefore more complicated. \subsection{Nonlinear Power Spectrum} The process of nonlinear gravitational clustering can be quantified statistically. Here I will only discuss the lowest-order statistical description given by the nonlinear power spectrum $P(k)$ of the matter density field. We have already discussed the linear power spectrum in Sec.~3.3. The triangles in Figure~\ref{fig:pow} show the nonlinear $P(k)$ computed from the particle positions in numerical simulations of three C+HDM models. The hierarchical nature of gravitational collapse in these models is illustrated by the fact that the high-$k$ modes have become strongly nonlinear whereas the low-$k$ modes are still following the linear power spectrum. The fact that the lowest several $k$ modes are still linear at $z=0$ ensures that our choice of the simulation box size (100 Mpc) is large enough to include all waves that have become nonlinear at present. The calculation of the fully evolved $P(k)$ for a given model is a laborious task involving the execution of high-resolution $N$-body simulations. Fortunately, some recent progress has been made in constructing analytical fitting formulas for a wide range of interesting models. The strategy is to examine the mapping between the linear and nonlinear $P(k)$ for a small, selected set of models with $N$-body data and then to extract systematic behavior for a wider range of parameters. The work carried out thus far has investigated scale-free models with a power-law power spectrum,~\cite{ham91,jain95} pure CDM and CDM with a cosmological constant $\Lambda$ (LCDM) models,~\cite{jain95,pd96,ma98} and C+HDM models.~\cite{ma98} The proposed formulas typically have a functional form that is motivated by analytical solutions in asymptotic regimes, but in order to obtain accurate approximations, the coefficients are calculated from fits to the nonlinear $P(k)$ computed from the numerical simulations. These approximations have provided physical insight into the process of nonlinear collapse and much practical convenience in incorporating the prominent nonlinear effects illustrated in Figure~\ref{fig:pow}. The formula applicable for the widest range of cosmological models thus far is given below. It maps the density variance $\Delta(k)\equiv 4\pi k^3P(k)$ in the linear and nonlinear regimes by~\cite{ma98} \begin{eqnarray} && {\Delta_{\rm nl}(k)\over \Delta_{\rm l}(k_0)} = G\left({\Delta_{\rm l}(k_0) \over g_0^{1.5}\,\sigma_8^\beta} \right) \,,\nonumber\\ && G(x)=[1+\ln(1+0.5\,x)]\,{1+0.02\,x^4 + c_1\,x^8/g^3 \over 1+c_2\,x^{7.5}}\,. \label{master} \end{eqnarray} Note that $\Delta_{\rm l}$ and $\Delta_{\rm nl}$ are evaluated at different wavenumbers, where $k_0=k\,(1+\Delta_{\rm nl})^{-1/3}$ corresponds to the pre-collapsed length scale of $k$. The parameter $\sigma_8$ is the rms linear mass fluctuation on $8\,h^{-1}$ Mpc scale at the redshift of interest, and $\beta=0.7+10\,\Omega_\nu^2$. The functions $g_0=g(\Omega_{\rm m},\Omega_\Lambda)$ and $g=g(\Omega_{\rm m}(a),\Omega_\Lambda(a))$ are, respectively, the relative growth factor for the linear density field at present day and at $a$ for a model with a present-day matter density $\Omega_{\rm m}$ and a cosmological constant $\Omega_\Lambda$. A good approximation is given by~\cite{lahav} \begin{equation} g ={5\over 2}\Omega_{\rm m}(a) [ \Omega_{\rm m}(a)^{4/7}-\Omega_\Lambda(a)+\left(1+ \Omega_{\rm m}(a)/2\right) \left(1+ \Omega_\Lambda(a)/70 \right)]^{-1}\,, \end{equation} and $\Omega_{\rm m}(a)=\Omega_{\rm m}\,a^{-1}/[1+\Omega_{\rm m}(a^{-1}-1) +\Omega_\Lambda(a^2-1)]$ and $\Omega_\Lambda(a)=\Omega_\Lambda\,a^2/[1+\Omega_{\rm m}(a^{-1}-1)+\Omega_\Lambda(a^2-1)]$. The time dependence is in factors $\sigma_8^\beta$ and $g$. For CDM and LCDM models, a good fit is given by $c_1=1.08\times 10^{-4}$ and $c_2=2.10\times 10^{-5}$. For C+HDM, a good fit is given by $c_1=3.16\times 10^{-3}$ and $c_2=3.49\times 10^{-4}$ for $\Omega_\nu=0.1$, and $c_1=6.96\times 10^{-3}$ and $c_2=4.39\times 10^{-4}$ for $\Omega_\nu=0.2$. \subsection{High-Redshift Constraints} The C+HDM models discussed thus far are a class of models bridging the much-studied albeit troubled pure CDM and pure HDM models. They are parameterized by the neutrino density parameter $0<\Omega_\nu< 1$, or equivalently, by the neutrino mass $0< m_\nu <93 h^2$ eV (see Eq.~(4)). The original motivation for examining these mixed models is to study whether the free-streaming effect introduced by the massive neutrinos could suppress the growth of density perturbations below the free-streaming scale, and thereby alleviate the problem of excess small-scale clustering in the standard CDM model. As illustrated in Figure~\ref{fig:pow}, massive neutrinos do indeed reduce the amplitude of clustering at large $k$, and a larger $\Omega_\nu$ leads to smaller high$-k$ power. Any viable cosmological model that provides a good statistical match to the local universe must also reproduce the appropriate evolutionary history out to high redshifts. Although gravitational clustering on galactic scales is indeed reduced in C+HDM models, providing a better match to low-redshift observations,\cite{ma96,chdm} this suppression in the clustering power at high redshifts has been found to pose serious problems for some models. Studies based on semi-analytic theories and dissipationless numerical simulations have shown~\cite{mb94b,dla} that flat C+HDM models with $\Omega_\nu>0.2$ do not produce enough early structure to explain the statistics of damped Ly$\alpha$ systems at redshift $z\ge 2$. More recent work~\cite{ma97} has included the effects of gas ionization and dissipation in the theoretical calculations and has compared the results to new data for damped Ly$\alpha$ systems at even higher redshift $z\sim 4$. It is found that ionization of hydrogen in the outskirts of halos and gaseous dissipation near the halo centers tend to exacerbate the problem of late galaxy formation. The amount of dense gas associated with the damped systems falls well below that observed, even for the flat $\Omega_\nu=0.2$ C+HDM model. This has placed an upper bound of $\sim 5$ eV on the sum of $\nu_e$, $\nu_\mu$, and $\nu_\tau$ masses, which is much more stringent than the upper limits given by current particle experiments (see Sec.~1). Let us briefly discuss the limitations and uncertainties in these calculations. Still debated is the nature of damped Ly$\alpha$ absorption -- whether it is due to intervening large, rapidly-rotating disk galaxies~\cite{pw97} with circular velocities $\go 200$ km s$^{-1}$, or infalling and randomly moving protogalactic gas clumps in dark matter halos~\cite{haeh98} with virial velocities of $\sim 100$ km s$^{-1}$. When kinematic considerations are included, the latter model may have problems balancing the high energy dissipation rate caused by cloud collisions.~\cite{mm98} This uncertainty aside, in either scenario, a host dark halo of velocity at least $\sim 100$ km s$^{-1}$ is needed to reproduce the large velocity widths and asymmetries of the observed low-ionization lines associated with Ly$\alpha$ systems. Uncertainties associated with the finite resolution of simulations have also been studied in some detail.~\cite{gardner97} It is found that even when the contribution from the numerically unresolved halos with velocities $v \lo 100$ km s$^{-1}$ is included, the absorption incidence is increased by at most a factor of 2. This is insufficient to erase the discrepancies reported for flat C+HDM models with $\Omega_\nu > 0.2$. The upper bound on neutrino masses from current cosmological studies is therefore $\sim 5$ eV. \section*{Acknowledgments} I thank the organizers P. Langacker and K. T. Mahanthappa of the TASI-98 School for their hospitality. This work was supported by the National Scalable Cluster Project at the University of Pennsylvania, the National Center for Supercomputing Applications, and a Penn Research Foundation Award. \section*{References}
\section{Introduction} In a recent paper \cite{Zhang} Zhang has proposed the idea that the components of the electron-pair condensate and those of the antiferromagnetic order-parameter form a five-component order parameter in an approximately SO(5)-invariant theory. Anisotropies small in comparison to the SO(5)-invariant interaction break this symmetry. They depend in particular on the chemical potential and thus on doping. This allows him to desribe the transition from antiferromagnetism in the half-filled system to superconductivity under moderately weak doping. This phenomenological picture has been tested by numerical calculations for small Hubbard-systems \cite{Hubbard,tJ-Hubbard} and $t$-$J$-models \cite{tJ,tJ-Hubbard}, for which good agreement with the predictions from Zhang's theory are obtained. It has also been shown that spin-ladders observe SO(5)-symmetry \cite{ladder1,ladder2} or even SO(8)-symmetry \cite{SO8}. Returning to higher-dimensional systems, it is pointed out here that Zhang's derivation of the effective interaction $V_{\rm eff}$ based on his Hamiltonian $H_{\rm a}$ seems to be not correct. (Actually we struggled for quite a while trying to understand his derivation). However, the orthogonality constraint introduced several pages after this 'derivation' gives the key to an effective interaction very similar to that given by Zhang. After having established this in sect. 2 it is shown in sect. 3, that the entropy contains squares of these constraint terms. Thus the orthogonality constraints are not strict requirements, but their fulfillment maximizes the entropy and lowers the free energy. Finally in sect. 4 it is argued, that also in the ground-state these constraints are likely to hold. \section{Effective Interaction and the Orthogonality Constraint} Zhang introduces a five-component order-parameter $n_i$, $i=1..5$. The operators associated with these components are given explicitely in the appendix. Two components, $n_1$ and $n_5$, describe the real part and the imaginary part of the superconducting condensate, the three other components $n_2$, $n_3$, and $n_4$, are the three Cartesian components of the staggered magnetization. He assumes that a strong interaction which may be described by a Ginzburg-Landau interaction, leads to a symmetry breaking below a critical temperature $T_c$ and has an SO(5) symmetry, thus allowing both superconducting and antiferromagnetic order. This leading interaction is a function of $\sum_i n_i^2$ only and does not give any preference to superconducting or antiferromagnetic order. In addition there is a weaker interaction $H_{\rm a}$ which introduces anisotropy into the system \begin{equation} H_{\rm a} = \frac{L^2_{1,5}}{2\chi_c} + \frac{L^2_{1,2}+L^2_{1,3}+L^2_{1,4}+L^2_{2,5}+L^2_{3,5} +L^2_{4,5}}{2\chi_{\pi}} + \frac{L^2_{2,3}+L^2_{2,4}+L^2_{3,4}}{2\chi_s} -\frac g2(n^2_2+n^2_3+n^2_4) -2\mu L_{1,5}. \end{equation} The $L_{a,b}$ are operators bilinear in the electron creation and annihilation operators. They are given explicitely in appendix A. $\mu$ is the chemical potential. The operators $L$ obey $L_{b,a}=-L_{a,b}$ and the commutator relations \begin{equation} [L_{a,b},L_{c,d}]=-i\delta_{b,c}L_{a,d}+i\delta_{a,c}L_{b,d} +i\delta_{b,d}L_{a,c}-i\delta_{a,d}L_{b,c} \label{comm} \end{equation} of the orthogonal group SO(5), moreover the vector $n$ is rotated by $L$, \begin{equation} [L_{a,b},n_c]=-i\delta_{b,c}n_a+i\delta_{a,c}n_c. \end{equation} The derivative terms of Zhang's interaction are left out here, since only the global state is investigated. The effective potential is determined as the minimum of the Hamiltonian for given order-parameter $n_1$, ... $n_5$, which Zhang normalizes to \begin{equation} \sum_i n_i^2 = 1. \label{norm} \end{equation} Since one considers the interaction for a macroscopic system, all the components of $L$ and $n$ can be considered as classical quantities and the obvious minimum is obtained for $L_{1,5}=2\mu\chi_c$, whereas all other components of $L$ vanish, which yields what I call $V^{\rm naiv}_{\rm eff}$ \begin{equation} V^{\rm naiv}_{\rm eff}(n) = -2\mu^2 \chi_c - \frac g2 (n^2_2+n^2_3+n^2_4). \end{equation} It would mean that the only effect of the chemical potential is a lowering of the energy of the system, but it would have no effect on an anisotropy in the order-parameter space. This obviously is at variance with Zhang's claim for the effective interaction \begin{equation} V^{\rm Zhang}_{\rm eff}(n) = -2\mu^2(n^2_1+n^2_5) [\chi_c(n^2_1+n^2_5)+\chi_{\pi}(n^2_2+n^2_3+n^2_4)] -\frac g2 (n^2_2+n^2_3+n^2_4). \end{equation} An indication that Zhang's Lagrangian \begin{eqnarray} {\cal L}_{\rm a}&=&\sum_{a<b} \frac{\chi_{ab}}2 \omega_{ab}^2 + V(n) \\ \omega_{ab}&=&n_a(\partial_{\tau}n_b-iB_{bc}n_c) -(a\rightarrow b) \label{omega} \end{eqnarray} is not equivalent to $H_{\rm a}$, is obvious from the fact that in Zhang's Lagrangian the only allowed rotations are in the plain spanned by $n$ and $\partial_{\tau}n - iBn$, whereas it does not permit a rotation around a second perpendicular plain (which exists in 5 dimensions) as it should. Thus the description by ${\cal L}_{\rm a}$ is incomplete. However, several pages later he introduces the constraints \begin{equation} \epsilon_{a,b,c,d,e}n_cL_{d,e}=0. \label{ortho} \end{equation} These constraints, (although not proven in Zhang's paper, since (\ref{omega}) is not granted) are the key for the effective interaction. If they are fulfilled, then one can conclude, that $L$ is of the form \begin{equation} L_{a,b} = c_a n_b -c_b n_a. \label{cn} \end{equation} (This can be found in the following way: If $n_1=1$, and all other $n_i$ vanish, then one finds from (\ref{ortho}) $L_{a,b}=0$, if both $a\not= 1$ and $b\not= 1$. However $L_{a,1}=-L_{1,a}=c_a$ for $a\not= 1$ can be chosen with arbitrary $c_a$. Thus one obtains in this special case eq. (\ref{cn}). The relation (\ref{cn}) is form-invariant under SO(5)-rotations. Thus it holds for general $n_i$.) Eq. (\ref{cn}) has two consequences: (i) One can now parametrize $H_{\rm a}$ in terms of the $c_i$ and determine the minimum as one varies the coefficients $c$. One finds the minimum for \begin{equation} c_2=c_3=c_4=0, \quad c_1=\frac{2\mu n_5}N, \quad c_5=-\frac{2\mu n_1}N \end{equation} with \begin{equation} N= \frac{n_1^2+n_5^2}{\chi_c} + \frac{n^2_2+n^2_3+n^2_4}{\chi_{\pi}}. \end{equation} {From} this one obtains the effective interaction \begin{equation} V_{\rm eff} = -2\mu^2 \frac{n_1^2+n_5^2}N -\frac g2 (n_2^2+n_3^2+n_4^2). \end{equation} For $\chi_c=\chi_{\pi}$ it agrees with $V^{\rm Zhang}_{\rm eff}$. If $\chi_c \not= \chi_{\pi}$, then the effective potentials are different (actually in first order in $\chi_c-\chi_{\pi}$ there is still agreement), but many conclusions Zhang has drawn will continue to hold. (ii) The constraints imply, that out of the representations $(\lambda_1,\lambda_2)$ the only allowed states have representations of the form $(\lambda_1,0)$. Let me explain this shortly. The irreducible representations of SO(5) are characterized by two (non negative) numbers $(\lambda_1,\lambda_2)$ with $\lambda_1 \geq \lambda_2 \geq 0$. The first number is the largest eigenvalue of one of the operators $L$, say $L_{1,2}$. Next consider the subspace of the states in this representation with this eigenvalue $\lambda_1$ for $L_{1,2}$. Then $\lambda_2$ is the largest eigenvalue of one of the operators $L$, that commutes with the first one, which may be $L_{3,4}$ in this subspace. Semiclassically, that is for large $\lambda_1$ and $\lambda_2$, one may replace the operators $L_{a,b}$ by their expectation values. Then the eigenvalues of the antisymmetric matrix $L$ are $\pm i\lambda_1$, $\pm i\lambda_2$ and 0. If $L$ has the form (\ref{cn}), then \begin{equation} \lambda_1 = \sqrt{\sum_i c_i^2 \sum_j n_j^2 - (\sum_i c_i n_i)^2}, \quad \lambda_2=0. \end{equation} Therefore the main problem left is to understand why the orthogonality constraints should hold. This will be done in the next section. \section{Entropy as Source of the Orthogonality Constraint} We claim that the orthogonality constraints are an effect of entropy. It will be shown that a system even without any interaction has an entropic parts of the form \begin{equation} -\big(\epsilon_{a,b,c,d,e}n_cL_{d,e}\big)^2 \end{equation} which make it favorable for the system to obey the orthogonality constraints. Thus (\ref{ortho}) is not a strict constraint, but if it is fulfilled, then the entropy assumes a maximum. \subsection{Entropy of the Heisenberg Antiferromagnet} First consider the Heisenberg antiferromagnet as an example, which Zhang has rightly mentioned in his paper. The Heisenberg antiferromagnet consists of two sublattices with magnetizations ${\bf m}_1$ and ${\bf m}_2$. We expand the entropy in powers of ${\bf m}_1^2$ and ${\bf m}_2^2$. Since the two sublattice-magnetizations describe the behaviour on different sublattices, the entropy is a sum of contributions on these sublattices \begin{equation} S=S_0-c_1({\bf m}_1^2+{\bf m}_2^2)-c_2\big(({\bf m}_1^2)^2 +({\bf m}_2^2)^2\big)-... \end{equation} Then denoting the homogeneous magnetization by ${\bf m}_0$ and the staggered magnetization by ${\bf m}_{q_0}$ we may write ${\bf m}_1={\bf m}_0+{\bf m}_{q_0}$ and ${\bf m}_2={\bf m}_0-{\bf m}_{q_0}$ and obtain the entropy \begin{equation} S=S_0-2c_1({\bf m}_0^2+{\bf m}_{q_0}^2) -2c_2\big({\bf m}_0^2+{\bf m}_{q_0}^2\big)^2 -8c_2\big({\bf m}_0\cdot{\bf m}_{q_0}\big)^2-... \end{equation} Thus if there is no coupling in the interaction between ${\bf m}_0$ and ${\bf m}_{q_0}$, the system prefers to have the homogeneous and the staggered magnetization orthogonal to each other due to the contribution $-8c_2({\bf m}_0\cdot{\bf m}_{q_0})^2$ in the entropy. Remember that the entropy enters into the free energy with a minus sign ($F=E-TS$), and thus ${\bf m}_0\cdot{\bf m}_{q_0}=0$ will yield the minimum of the free energy. Thus there is not a strict constraint ${\bf m}_0\cdot{\bf m}_{q_0}=0$ for the Heisenberg antiferromagnet, but there is a term proportional to the square $({\bf m}_0\cdot{\bf m}_{q_0})^2$ in the free energy, which favors the constraint to be obeyed. \subsection{The entropy} It will now be shown that the entropy of the more general SO(5)-invariant system contains terms of type $(\epsilon_{a,b,c,d,e}n_cL_{d,e})^2$. We will actually enlarge the system of operators $L_{a,b}$ to those of an SO(8) (for details see in the appendix). To determine the entropy we start from the Hamiltonian \begin{equation} H_{\Omega} = \sum_{a,b} \Omega_{a,b} L_{a,b}, \quad \Omega_{b,a}=-\Omega_{a,b}. \label{HOmega} \end{equation} where the set of our $L_{a,b}$ also includes the $n_a=L_{a,0}$. We do no longer use the normalization (\ref{norm}). We note, that recently a U(4)-scheme including the operators $L_{a,b}$ with $a,b=0..5$ and $L_{6,7}$ and their sub-groups has been considered in \cite{U4} (note that SU(4) is isomorphous to SO(6)). The $\Omega$ are introduced as Lagrange multipliers and will be adjusted to yield given expectation values of $L_{a,b}$, and the entropy will be calculated up to fourth order in $L$. $\Omega$ is an eight-dimensional antisymmetric real matrix. Since the matrix is antihermitean, its eigenvalues are purely imaginary and occur pairwise. Thus the eigenvalue equation can be written \begin{equation} \sum_b \Omega_{a,b} (x^{(k)}_b \pm i y^{(k)}_b) = \pm i \omega^{(k)} (x^{(k)}_a \pm i y^{(k)}_a) \end{equation} with $k=1..4$ and real vectors $x^{(k)}$ and $y^{(k)}$. The vectors $x^{(k)}$ and $y^{(k)}$ are orthogonal to each other. If they are normalized, then $\Omega$ may be represented \begin{equation} \Omega_{a,b} = \sum_k \omega^{(k)} (x^{(k)}_a y^{(k)}_b - x^{(k)}_b y^{(k)}_a). \end{equation} Next we perform a special orthogonal transformation, so that $x^{(k)}$ and $y^{(k)}$ are oriented in appropriate directions, e.g. so that $x^{(1)}$, $y^{(1)}$ point in the 5- and 1- direction, $x^{(2)}$, $y^{(2)}$ in the 2- and 3-direction, $x^{(3)}$ and $y^{(3)}$ in the 0- and 7-direction and $x^{(4)}$ and $y^{(4)}$ in the 4- and 6- direction, resp. After this special orthogonal transformation $H$ reads \begin{eqnarray} H_{\rm trans} &=& 2(\omega^{(1)} N^0_{0,1,1} + \omega^{(2)} N^0_{0,1,\sigma^z} +\omega^{(3)} N^0_{0,g,1} -\omega^{(4)} N^0_{0,g,\sigma^z}) \nonumber \\ &=& \sum_{k,s} ( \omega^{(1)} +\omega^{(2)} s + \omega^{(3)} g(k) -\omega^{(4)} g(k) s) (c^{\dagger}_{k,s} c_{k,s} - \frac 12). \label{Htrans} \end{eqnarray} (with $s=\pm 1$, $g(k)=\pm 1$, $g(k+q_0)=-g(k)$). Thus for momenta $k$ and $k+q_0$ one has in total $2^4$ states constructed out of four single-particle states, which may be either occupied or unoccupied and which contribute the energies $\pm \frac 12 \epsilon_{s,g}$ depending on whether the state is occupied or unoccupied \begin{equation} \epsilon_{s,g} = \omega^{(1)} + \omega^{(2)} s + \omega^{(3)} g - \omega^{(4)} gs. \end{equation} Then we obtain for the partition function $Z$ \begin{equation} \ln Z=\sum \ln(\exp(\frac 12\beta\epsilon) +\exp(-\frac 12\beta\epsilon)) =\sum (\ln 2 + \frac{x^2}2 - \frac{x^4}{12} + O(x^6)), \quad x=\frac{\beta\epsilon}2. \end{equation} Summation over the four states yields \begin{eqnarray} \sum x^2 &=& \beta^2 \sum_i \omega^{(i)2} \\ \sum x^4 &=& \frac{\beta^4}4 (\sum_i \omega^{(i)4} +6 \sum_{i<j} \omega^{(i)2} \omega^{(j)2} - 24 \omega^{(1)} \omega^{(2)} \omega^{(3)} \omega^{(4)} ). \end{eqnarray} Similarly we expand the entropy $S$ \begin{equation} S=k_B \sum (\ln 2 - \frac{x^2}2 + \frac{x^4}4 + O(x^6)) \end{equation} and determine the expectation value of the quantity $L^{(i)}$ conjugate to $\omega^{(i)}$ (only the first component is given; the others are obtained by permutation), \begin{equation} L^{(1)} = \frac{\partial \ln Z}{2\beta\partial \omega^{(1)}} = \frac{\beta}2 \omega^{(1)} -\frac{\beta^3}{24}\big(\omega^{(1)3} + 3\omega^{(1)}(\omega^{(2)2}+\omega^{(3)2}+\omega^{(4)2}) -6\omega^{(2)}\omega^{(3)}\omega^{(4)} \big) + O(\omega^5) \end{equation} and express $\omega^{(i)}$ in terms of $L^{(i)}$, \begin{equation} \beta\omega^{(1)} = 2L^{(1)} + \frac 23\big( L^{(1)3} +3L^{(1)}(L^{(2)2}+L^{(3)2}+L^{(4)2}) -6L^{(2)}L^{(3)}L^{(4)} \big) +O(L^5). \end{equation} In diagonal form one has $L^{(1)}=L_{5,1}$, etc. Then $S$ for the subspace of the electrons with momenta $k$ and $k+q_0$ reads \begin{equation} S/k_B=4\ln2-2\sum_i L^{(i)2} -\frac 13 \big(\sum_i L^{(i)4} +6\sum_{i<j} L^{(i)2} L^{(j)2} -24L^{(1)}L^{(2)}L^{(3)}L^{(4)}\big) +O(L^6). \label{SLeig} \end{equation} Thus we have expressed the entropy in terms of the eigenvalues $\pm i L^{(i)}$ of the $8\times8$-matrix $(L_{a,b})$ of the expectation values $L_{a,b}$. We will express it now by the matrix-elements of $L$. First we have \begin{eqnarray} \sum_i L^{(i)2} &=& -\frac 12 {\rm tr} L^2 = -\sum_{a<b} L_{a,b}^2, \\ \sum_i L^{(i)4} &=& \frac 12 {\rm tr} L^4 = \sum_{a<b} L_{a,b}^4 + 2 \sum_{a,b<c} L_{a,b}^2 L_{a,c}^2 + 4\hspace{-5mm} \sum_{a<b,a<c<d,b\not=c,b\not=d} \hspace{-11mm} L_{a,c}L_{a,d}L_{b,c}L_{b,d}, \\ \hspace{-2mm} \sum_i L^{(i)4} + 2\sum_{i<j} L^{(i)2} L^{(j)2} &=& \frac 14 ({\rm tr} L^2)^2 = \sum_{a<b} L_{a,b}^4 + 2\sum_{a,b<c} L_{a,b}^2 L_{a,c}^2 +2\hspace{-3mm} \sum_{a<b,a<c<d,b\not=c,b\not=d} \hspace{-6mm} L_{a,b}^2 L_{c,d}^2, \\ L^{(1)}L^{(2)}L^{(3)}L^{(4)} &=& {\rm pf}(L), \end{eqnarray} where ${\rm pf}(L)$ is the Pfaffian of $L$. In our case the indices $a$, $b$ of the elements $L_{a,b}$ are numbered from 0 to 7. Then the Pfaffian (which is defined only for antisymmetric matrices) is the sum $\sum_{k=1}^7 (-)^{k-1} L_{0,k} P_k$, where $P_k$ is the Pfaffian of the matrix obtained from the matrix $L$ by deleting the rows and columns with index 0 (that is the first one) and index $k$. One continues recursively until the Pfaffian of a matrix with no entry is left, which is defined to equal 1. (If one starts with a matrix of odd dimension, then finally one arrives at the Pfaffian of the $1\times1$ matrix with entry 0, since it has to be antisymmetric. This Pfaffian is defined to equal 0. Therefore the Pfaffian of a matrix of odd dimensions vanishes.) We mention that the determinant of an antisymmetric matrix equals the square of its Pfaffian. For $2\times2$ and $4\times4$ matrices the Pfaffians read \begin{eqnarray} {\rm pf}\left(\begin{array}{cc} 0 & L_{a,b} \\ -L_{a,b} & 0 \end{array} \right) &=& L_{a,b} \\ p_{a,b,c,d}:= {\rm pf} \left(\begin{array}{cccc} 0 & L_{a,b} & L_{a,c} & L_{a,d} \\ -L_{a,b} & 0 & L_{b,c} & L_{b,d} \\ -L_{a,c} & -L_{b,c} & 0 & L_{c,d} \\ -L_{a,d} & -L_{b,d} & -L_{c,d} & 0 \end{array}\right) &=& L_{a,b} L_{c,d} - L_{a,c} L_{b,d} + L_{a,d} L_{b,c}. \end{eqnarray} Then we have \begin{equation} S/k_B=4\ln 2 + {\rm tr} L^2 -\frac 13\big(\frac 34 ({\rm tr} L^2)^2 -{\rm tr} L^4 -24 {\rm pf}(L) \big). \end{equation} (Note that ${\rm tr} L^2$ is negative). Further algebraic manipulations yield \begin{eqnarray} -{\rm tr} L^4 &=& -\frac 12 ({\rm tr} L^2)^2 + 4\sum_{a<b<c<d} p_{a,b,c,d}^2, \\ 24{\rm pf}(L) &=& 4\sum_{a<b<c<d,\, e<f<g<h} \epsilon_{a,b,c,d,e,f,g,h} \,p_{a,b,c,d} \,p_{e,f,g,h}. \end{eqnarray} Putting all the contributions together we obtain \begin{eqnarray} S/k_B &=& 4\ln2 + {\rm tr} L^2 -\frac 1{12} ({\rm tr} L^2)^2 \nonumber \\ &-&\frac 43 \sum_{0<a<b<c} \big( p_{0,a,b,c} - \sum_{d<e<f<g} \epsilon_{a,b,c,d,e,f,g} \, p_{d,e,f,g} \big)^2 + O(L^6). \end{eqnarray} This corresponds to the separation of (\ref{SLeig}) into \begin{eqnarray} \hspace{-5mm}S/k_B &=& 4 \ln 2 -2 \sum_i L^{(i)2} - \frac 13(\sum_i L^{(i)2})^2 \nonumber \\ &-&\frac 43 \big( (L^{(1)}L^{(2)}-L^{(3)}L^{(4)})^2 + (L^{(1)}L^{(3)}-L^{(2)}L^{(4)})^2 + (L^{(1)}L^{(4)}-L^{(2)}L^{(3)})^2 \big) +O(L^6). \label{SLeig2} \end{eqnarray} Thus the entropy consists of a completely rotational invariant contribution depending only on ${\rm tr} L^2$ and a negative sum of squares of $p_{0,a,b,c} \pm p_{d,e,f,g}$. Note that the sum over $d,e,f,g$ contains exactly one term $\pm p$. Thus a maximum of the entropy is reached, when the arguments of all the squares vanish. Provided $L_{a,b}$ vanishes if $a$ or $b$ equals 6 or 7 as assumed in the SO(5) theory, then out of the 35 squares 20 vanish identically, 10 have the form $\epsilon_{a,b,c,d,e} n_c L_{d,e}$, which when required to vanish are Zhang's orthogonality constraints, and 5 have the form $\epsilon_{a,b,c,d,e}L_{b,c}L_{d,e}$. One can easily see from (\ref{cn}), that if Zhang's orthogonality constraints are fulfilled, then also these latter quantities vanish. Thus the orthogonality constraints are not a strict requirements, but their fulfillment lowers the free energy. We mention, that the vanishing of the squares in the second line of (\ref{SLeig2}) implies two types of solutions, either one $L^{(i)}$ can be different from zero and the other $L^{(j)}$ vanish, or all $|L^{(i)}|$ are equal, and the product of the four $L^{(i)}$ is positive. At half-filling (no doping) one has $L_{1,5}=0$. Then in the antiferromagnetic state, that is for non-vanishing components $L_{2,0}$, $L_{3,0}$, and $L_{4,0}$, but otherwise vanishing components $L$, only one eigenvalue $L^{(i)}$ is different from zero. As soon as one has some doping $L_{1,5}$ is different from 0. If the system is still antiferromagnetic, then more than one eigenvalue differs from zero. On the other hand, if the modes associated with $n_6=L_{6,0}$ or $n_7=L_{7,0}$ are massive, as assumed, then it costs energy to have all four eigenvalues $L^{(i)}$ different from zero. Then it becomes preferable to have the superconductiong components $n_1=L_{1,0}$ and $n_5=L_{5,0}$ different from zero and to have vanishing antiferromagnetism, since then again one has only one nonvanishing eigenvalue $L^{(i)}$. \section{Limits for the ground state} The entropy argument given in the preceeding section can be applied in the vicinity of the critical temperature. At low temperatures the contribution of the entropy to the free energy decreases, since it enters with the factor $T$. Therefore we consider separately the situation at low temperatures. \subsection{Antiferromagnet} Let us start with a simple consideration for the antiferromagnet. Assume again two sublattices with magnetization ${\bf m}_1$ and ${\bf m}_2$. Assume they are restricted by upper bounds $|{\bf m}_1|\le 1$, $|{\bf m}_2|\le1$. Suppose now the system has a homogeneous magnetization ${\bf m}_0$. If now a staggered magnetization parallel to ${\bf m}_0$ is added, then apparently $|{\bf m}_{q_0}|\le 1-|{\bf m}_0|$. If however the staggered magnetization is perpendicular to the homogeneous one, then one has the weaker restriction $|{\bf m}_{q_0}|\le \sqrt{1-{\bf m}_0^2}$. Thus the antiferromagnetic interaction can act more strongly, if the staggered magnetization is perpendicular to the homogeneous one. Thus again ${\bf m}_0\cdot{\bf m}_{q_0}=0$ is fulfilled. \subsection{Hartree-Fock-Bogoliubov ground state} We have seen, that at low temperatures the bounds on the appropriate quantities (order-parameters) are important for the ordering of the ground-state. Therefore we will consider the bounds of the eigenvalues $\pm i L^{(i)}$ of the matrix $L$. If for a fixed matrix $\Omega$ in (\ref{HOmega}) one takes the low temperature limit $\beta\rightarrow\infty$, then normally a pure Hartree-Fock-Bogoliubov state instead of a mixed state remains. Depending on the sign of $\epsilon_{s,g}$ the occupation number $n_{s,g}$ assumes one of its extremal values 0 and 1 in the diagonal representation $H_{\rm trans}$ (\ref{Htrans}) \begin{equation} n_{s,g} = \frac 12 -\frac 12 {\rm sign} \epsilon_{s,g}. \end{equation} {From} this we obtain the eigenvalues $L^{(i)}$ \begin{eqnarray} L^{(1)} &=& \frac 12 \sum_{s,g} n_{s,g} - 1 \\ L^{(2)} &=& \frac 12 \sum_{s,g} s\, n_{s,g} \\ L^{(3)} &=& \frac 12 \sum_{s,g} g\, n_{s,g} \\ L^{(4)} &=& \frac 12 \sum_{s,g} sg\, n_{s,g}. \end{eqnarray} Putting now $n_{s,g}=0$ or 1 in all combinations one finds two types of solutions for $L^{(i)}$. In the first class one $L^{(i)}=\pm 1$ and the other $L^{(j)}=0$, whereas in the second class one has $L^{(i)}=\pm \frac 12$ for all $i$ with the restriction for the signs, that the product of all $L^{(i)}$ is positive. One easily realizes that in all these cases the squares in the second line of eq. (\ref{SLeig2}) vanish. As a consequence the orthogonality constraints are fulfilled for these states. The eigenvalues obtained for the Hartree-Fock-Bogoliubov ground-state are the extremal ones. For correlated states these eigenvalues can only be reduced. More precisely, the range of the set of possible eigenvalues $L^{(1)}$, ... $L^{(4)}$ lies in the convex volume bounded by the extremes given above. For an interaction quadratic in the operators $L$ the energy assumes extremal eigenvalues for extremal eigenvalues $L$ in the case of symmetry breaking. Without symmetry breaking the eigenvalues would vanish. We have argued before in favour of only one non-vanishing eigenvalue $L^{(i)}$. Reduced to SO(5) this would be $\lambda_1$, whereas the second largest (vanishing one) is $\lambda_2$ of the representation $(\lambda_1,\lambda_2)$. These correspond to the representations ($\lambda_1$,0) actually found in the numerical calculations \cite{tJ,tJ-Hubbard}. \section{Conclusion} We have shown that the orthogonality constraints introduced by Zhang in his SO(5)-theory of high$T_c$-superconductivity play an important role in the mechanism for the transition from antiferromagnetism to superconductivity as a function of doping. We have further shown that these constraints are not strictly fulfilled, but that their fulfillment yields a maximum of the entropy (for fixed $\sum L_{a,b}^2$). At low temperatures the entropy plays a weaker role. However, if the interaction drives the ground-state to the case of extremal eigenvalues $L^{(i)}$, which may be very well the case for an interaction bilinear in the operators $L$, then again the orthogonality constraints are fulfilled. We have expanded our scheme to a (mathematically natural) SO(8)-scheme. In this scheme two types of solutions for maximal entropy or extremal eigenvalues appear. Since the added degrees of freedom are probably massive, only those solutions, for which one eigenvalue is different from zero and the other ones vanish yield the minimum in the free energy. These considerations did not take the microscopic interaction seriously into account. Work in this direction has to be done. {\bf Acknowledgment} I am indebted to Andreas Mielke and J\"urgen Stein for useful comments. \begin{appendix} \section{Operators in an SO(8) space} \subsection{Bilinear operators} All the operators in the SO(5) theory are of the form \begin{eqnarray} N^+_{q,f,\gamma}&=&\frac 12 \sum_{k,s,t} f(k) c^{\dagger}_{k+q,s} (\gamma \sigma^y)_{s,t} c^{\dagger}_{-k,t}, \\ N^0_{q,f,\gamma}&=&\frac 12 \sum_{k,s,t} f(k) \gamma_{s,t} (c^{\dagger}_{k+q,s} c_{k,t} - \frac 12\delta_{q,0}\delta_{s,t}), \\ N^-_{q,f,\gamma}&=&\frac 12 \sum_{k,s,t} f(k) c_{k+q,s} (\sigma^y \gamma)_{s,t} c_{-k,t}. \end{eqnarray} Here the summation $k$ runs over the Brillouin zone. The vector $q$ can either be 0 or $q_0$, where $2q_0$ is a reciprocal lattice vector. The staggered magnetization is described by the wave-vector $q_0$ as above. The function $f(k)$ stands either for 1 or $g(k)$ with $g(k)=g(-k)=-g(k+q_0)=\pm 1$. Finally $\gamma$ is a hermitean two by two matrix. It may be either the unit matrix or one of the Pauli matrices. First we consider the symmetry of $N^+$. If we exchange the two $c^{\dagger}$-operators, then we obtain \begin{equation} N^+_{q,f,\gamma} = N^+_{q,f(.+q),\sigma^y\gamma^T\sigma^y}. \end{equation} One has \begin{equation} f(.+q)=s_{q,f}f \mbox{ with } s_{q,f}=\left\{ \begin{array}{c} 1 \mbox{ for } q=0 \mbox{ or } f=1, \\ -1 \mbox{ for } q=q_0 \mbox{ and } f=g. \end{array} \right. \end{equation} and \begin{equation} \sigma^y\gamma^T\sigma^y = s_{\gamma} \gamma \mbox{ with } s_{\gamma} = \left\{ \begin{array}{c} +1 \mbox{ for } \gamma=1 \\ -1 \mbox{ for Pauli matrices } \end{array} \right. \end{equation} {From} this we conclude that only the six operators $N^+_{0,1,1}$, $N^+_{0,g,1}$, $N^+_{q_0,1,1}$, $N^+_{q_0,g,\sigma_\alpha}$ with $s_{q,f}s_{\gamma}=1$ are different from zero. The same holds for $N^-$. The hermitean adjoint operators are \begin{eqnarray} (N^-_{q,f,\gamma})^{\dagger} &=& s_{q,f} N^+_{q,f,\gamma^{\dagger}}, \\ (N^0_{q,f,\gamma})^{\dagger} &=& s_{q,f} N^0_{q,f,\gamma^{\dagger}}. \end{eqnarray} We obtain for the commutators \begin{eqnarray} {[}N^+_{q,f,\gamma},N^+_{q',f',\gamma'}]&=&0, \\ {[}N^-_{q,f,\gamma},N^-_{q',f',\gamma'}]&=&0, \\ {[}N^-_{q,f,\gamma},N^+_{q',f',\gamma'}] &=&-2N^0_{q'-q,ff'(.-q),\gamma\gamma'}, \\ {[}N^-_{q,f,\gamma},N^0_{q',f',\gamma'}] &=&N^-_{q-q',f(.-q')f',\gamma\gamma'}, \\ {[}N^0_{q,f,\gamma},N^+_{q',f',\gamma'}] &=&N^+_{q+q',f(.+q')f',\gamma\gamma'}, \\ {[}N^0_{q,f,\gamma},N^0_{q',f',\gamma'}] &=&\frac 12(N^0_{q+q',f(.+q')f',\gamma\gamma'} -N^0_{q+q',ff'(.+q),\gamma'\gamma}). \end{eqnarray} \subsection{Zhang's Operators} The operators introduced by Zhang, which obey the commutator relations (\ref{comm}) of an SO-group were \begin{eqnarray} L_{5,1}=&Q&=N^0_{0,1,1}, \\ L_{1+\alpha,1}=\frac 12 (\pi^{\dagger}_{\alpha}+\pi_{\alpha}), && \pi^{\dagger}_{\alpha}=N^+_{q_0,g,\sigma^{\alpha}}, \\ L_{5,1+\alpha}=\frac {-i}2 (\pi^{\dagger}_{\alpha}-\pi_{\alpha}), && \pi_{\alpha}=-N^-_{q_0,g,\sigma^{\alpha}}, \\ L_{1+\alpha,1+\beta} =\epsilon_{\alpha,\beta,\gamma}S_{0,\gamma}, && S_{0,\gamma}=N^0_{0,1,\sigma^{\gamma}}, \\ n_1=L_{1,0}=\frac 12(\Delta^{\dagger}+\Delta), && \Delta^{\dagger}=iN^+_{0,g,1},\\ n_5=L_{5,0}=\frac i2(\Delta^{\dagger}-\Delta), && \Delta=-iN^-_{0,g,1},\\ n_{1+\alpha}=L_{1+\alpha,0}&=&S_{q_0,\alpha} =N^0_{q_0,1,\sigma^{\alpha}}. \end{eqnarray} Here we have added $n_a=L_{a,0}$ to the group, since they obey the same commutator relations. \subsection{Extension to operators obeying an SO(8) group} Obviously one can expand the range of operators $L_{a,b}$ by including the other operators $N$ introduced above so that any pairs of particles, of holes or particle-hole pairs with total momentum 0 or $q_0$ in the singlet and triplet channel appear. This allows us to introduce components $L_{a,b}$ with $a,b=6$ or 7. This expansion does not imply, that the symmetry group for the phase transition will be enlarged. Actually one expects that the new components of the 'order-parameter' stay massive at the transition and will not contribute directly to the symmetry-breaking. It is wellknown \cite{Lieb, Yang} that there are additional operators adding or removing two electrons, $\eta^{\dagger}$ and $\eta$, which commute with the Hubbard Hamiltonian, which allows us by means of the commutator relations (\ref{comm}) to introduce \begin{eqnarray} L_{1,6}=\frac i2(\eta-\eta^{\dagger}), && \eta^{\dagger}=N^+_{q_0,1,1} \\ L_{5,6}=\frac 12(\eta^{\dagger}+\eta), && \eta=N^-_{q_0,1,1}, \\ L_{1+\alpha,6}&=&-N^0_{0,g,\sigma^{\alpha}}, \\ n_6=L_{6,0}&=&iN^0_{q_0,g,1}. \end{eqnarray} Finally there are 7 more operators left, which fulfill the appropriate commutator relations with \begin{eqnarray} L_{1,7}=\frac 12(\tilde{\eta}^{\dagger}+\tilde{\eta}), && \tilde{\eta}^{\dagger}=N^+_{0,1,1}, \\ L_{5,7}=\frac i2(\tilde{\eta}^{\dagger}-\tilde{\eta}), && \tilde{\eta}=N^-_{0,1,1}, \\ L_{1+\alpha,7} &=& iN^0_{q_0,g,\sigma^{\alpha}}, \\ L_{6,7} &=& -N^0_{q_0,1,1}, \\ n_7=L_{7,0} &=& -N^0_{0,g,1}. \end{eqnarray} \end{appendix}
\section{\@startsection{section}{1}{\z@}{-3.25ex plus -1ex minus -.2ex}{1.5ex plus .2ex}{\normalfont\bfseries}} \def\subsection{\@startsection{subsection}{2}{\z@}{-3.25ex plus -1ex minus -.2ex}{1.5ex plus .2ex}{\normalfont\itshape}} \@addtoreset{equation}{section} \renewcommand{\theequation}{\arabic{section}.\arabic{equation}} \def\ps@mathphys{\addtolength{\headheight}{5pt} \addtolength{\topmargin}{-15pt} \addtolength{\headsep}{15pt} \def\@oddhead{\hfil\begin{tabular}{r} {\tt math-ph/9904009} \\ CPT-99/P.3809 \end{tabular}} \let\@evenhead\@oddhead \def\@oddfoot{\hfil\thepage\hfil}\let\@evenfoot\@oddfoot} \makeatother \font\twelvebb=msbm10 scaled 1200 \newfam\bbfam \textfont\bbfam=\twelvebb \def\fam\bbfam{\fam\bbfam} \begin{document} \thispagestyle{mathphys} \begin{center} {\baselineskip 20pt {\large\bf On the Connes--Moscovici Hopf algebra associated to the diffeomorphisms of a manifold} \vskip 3ex {\sc Raimar Wulkenhaar} } \vskip 1ex {\it Centre de Physique Th\'eorique\\ CNRS-Luminy, Case 907\\ 13288 Marseille Cedex 9, France} \end{center} \vskip 3ex \begin{abstract} For our own education, we reconstruct the Hopf algebra of Connes and Moscovici obtained by the action of vector fields on a crossed product of functions by diffeomorphisms. We extend the realization of that Hopf algebra in terms of rooted trees as given by Connes and Kreimer from dimension one to arbitrary dimension of the manifold. In principle there is no modification, but in higher dimension one has to be careful with the order of cuts. The order problem leads us to speculate that in quantum field theory the sum of Feynman graphs which corresponds to an element of the Connes--Moscovici Hopf algebra could have a larger symmetry than the individual graphs. \\[1ex] \tabcolsep 0pt \begin{tabular}{lll} 1991 MSC:~~{} & 16W30~{} & Coalgebras, bialgebras, Hopf algebras \\ & 57R25 & Vector fields, frame fields \\ & 57R50 & Diffeomorphisms \\ & 81T15 & Perturbative methods of renormalization \end{tabular} \end{abstract} \section{Introduction} Recently two useful Hopf algebras were discovered -- by Alain Connes and Henri Moscovici in noncommutative geometry \cite{cm2} and by Dirk Kreimer in quantum field theory \cite{k1}. Connes and Kreimer showed that both Hopf algebras are intimately related \cite{ck}, via the language of rooted trees. Recently it was pointed out \cite{b} that the same algebra of rooted trees appears in numerical analysis. We refer to \cite{ck2} for a review of all these ideas. For a physicist, the Hopf algebra ${\cal H}_K$ of Kreimer \cite{k1} is not difficult to understand. The idea is to look at the divergence structure of Feynman graphs. There is a natural splitting of a Feynman graph $\gamma$ with non-overlapping subdivergences into two, given by a product of selected subdivergences $\gamma_1\cdots \gamma_n$ and the graph $\gamma\setminus(\gamma_1\cdots \gamma_n)$ left over from $\gamma$ by shrinking all $\gamma_i$, $i=1,\dots,n$, to a point. This is a standard operation in renormalization theory. Summing over all possibilities gives a coproduct operation on the algebra of polynomials in Feynman graphs. The unique antipode reproduces precisely the combinatorics of renormalization, i.e.\ it produces the local counterterms to make the integral corresponding to the divergent Feynman graph finite. The aim of this note is to review in some detail the construction of the Hopf algebra found by Connes and Moscovici, in order to ease its access for physicists interested in the Hopf algebra ${\cal H}_K$ of renormalization. The construction requires only some basic knowledge of classical differential geometry, which can be found in many books on this topic, for instance in \cite{cd}. More precisely, one needs the vertical and horizontal vector fields $Y$ and $X$ on the frame bundle over an oriented manifold and their transformation behavior under diffeomorphisms, as well as some familiarity with the push-forward and pull-back operations. These requisites are derived in section \ref{general}. New is the application of these vector fields to the crossed product, see section \ref{cross}, which defines the coproduct of $X,Y$ and leads to an additional operator $\delta$ on the crossed product. The operators $X,Y,\delta$ generate a Lie algebra. Its enveloping algebra ${\cal H}$ is a bialgebra with the coproduct obtained before, and there exists an antipode making it to a Hopf algebra, see section \ref{Lie}. The final section is devoted to the transformation of the commutative Hopf subalgebra ${\cal H}_{CM}$ of ${\cal H}$ into the language of rooted trees so that we can compare it with ${\cal H}_K$. We generalize the rooted trees given in \cite{ck} from dimension 1 to arbitrary dimension of the manifold. This generalization is quite obvious, but it has several consequences which are not visible in dimension 1. An element of ${\cal H}_{CM}$ is a sum of decorated planar rooted trees. The root is decorated by three spacetime indices necessary to describe parallel transport whereas the other vertices are decorated by a single spacetime index. This is closer to quantum field theory, where the decoration is given by primitive Feynman graphs without subdivergences. Interestingly, elements of ${\cal H}_{CM}$ are invariant under permutations of the decorations, whereas the individual trees representing Feynman graphs are not. This raises the question whether the sum of Feynman graphs which corresponds to an element of ${\cal H}_{CM}$ has a meaning in QFT. \section{The geometry of the frame bundle} \label{general} In this section we are going to derive in some detail the following well-known results on the principal fibre bundle $F^+$ of oriented frames on an $n$-dimensional manifold $M$: \begin{Proposition} Let $\{x^\mu\}_{\mu=1\dots,n}$ be the coordinates of $x\in M$ within a local chart of $M$ and $\{y^\mu_i\}_{\mu,i = 1,\dots n}$ be the coordinates of $n$ linearly independent vectors of the tangent space $T_xM$ with respect to the basis $\partial_\mu$. On $F^+$ there exist the following geometrical objects, written in terms of the local coordinates $(x^\mu,y^\mu_i)$ of $p \in F^+$: \begin{itemize} \item[(1)] an ${\fam\bbfam R}^n$-valued (soldering) $1$-form $\alpha$ with $\alpha^i = (y^{-1})^i_\mu d x^\mu~,$ \item[(2)] a $gl(n)$-valued (connection) $1$-form $\omega$ with $\omega^i_j = (y^{-1})^i_\mu ( d y^\mu_j + \Gamma^\mu_{\alpha\beta}\, y^\alpha_j dx^\beta)$, where $\Gamma^\mu_{\alpha\beta}$ depends only on $x^\nu~,$ \item[(3)] $n^2$ vertical vector fields $Y^i_j = y^\mu_j \partial^i_\mu~,$ \item[(4)] $n$ horizontal (with respect to $\omega$) vector fields $X_i = y^\mu_i ( \partial_\mu - \Gamma^\nu_{\alpha\mu} y^\alpha_j \partial^j_\nu)~.$ \end{itemize} A local diffeomorphism $\psi$ of $M$ has a lift $\tilde{\psi}: (x^\mu,y^\mu_i) \mapsto (\psi(x)^\mu,\partial_\nu \psi(x)^\mu y^\nu_i)$ to the frame bundle and induces the following transformations of the previous geometrical objects: \begin{itemize} \item[(1)] $(\tilde{\psi}^* \alpha)\big|_p = \alpha\big|_p~.$ \item[(2)] $(\tilde{\psi}^* \omega)\big|_p = (y^{-1})^i_\mu ( d y^\mu_j + \tilde{\Gamma}^\mu_{\alpha\beta}\, y^\alpha_j dx^\beta)$ is again a connection form, with \\[0.5ex] $\tilde{\Gamma}^\mu_{\alpha\beta}\big|_x = ((\partial\psi(x))^{-1})^\mu_\gamma \, \Gamma^\gamma_{\delta\epsilon}\big|_{\psi(x)} \, \partial_\alpha\psi(x)^\delta \partial_\beta \psi(x)^\epsilon + ((\partial\psi(x))^{-1})^\mu_\gamma \, \partial_\beta \partial_\alpha \psi(x)^\gamma ~,$ \item[(3)] $(\tilde{\psi}_* Y^j_i)\big|_p = Y^j_i\big|_p~,$ \item[(4)] $(\tilde{\psi}^{-1}_* X_i)\big|_p = y^\mu_i ( \partial_\mu - \tilde{\Gamma}^\nu_{\alpha\mu} y^\alpha_j \partial^j_\nu)$ is horizontal to $\tilde{\psi}^* \omega~.$ \end{itemize} \label{prpgen} \end{Proposition} The reader familiar with these notations can pass immediately to section \ref{cross} on page~\pageref{cross}. \subsection{Frame bundle} Let $M$ be an $n$-dimensional smooth and oriented manifold. We are going to consider the frame bundle $F^+$ over $M$ defined as follows. Let $T_xM$ be the tangent space at a given point $x \in M$. It is an $n$-dimensional vector space containing the tangent vectors at $x$ of curves in $M$ through $x$. A base in $T_xM$ is given by the $n$ tangent vectors $\partial_\mu:=\frac{\partial}{\partial x^\mu}$ of the coordinate lines in $M$. If $x$ has (in a given chart of its neighbourhood) the coordinates $\{x^\mu\} \equiv(x^1,\dots,x^n)$, we compute the tangent vector to a curve $C (t) = \{x^\mu(t)\}$: \begin{equation} \frac{d\phi(C(t))}{dt}\Big|_{t=0} =\frac{d x^\mu}{dt}\Big|_{t=0} \; \frac{\partial}{\partial x^\mu} \phi \Big|_x~, \end{equation} where $\phi:M \to {\fam\bbfam R}$ (or ${\fam\bbfam C}$) is an arbitrary function on $M$. According to Einstein's sum convention summation over pairs of identical upper and lower indices is self-understood. An arbitrary vector $Y_j \in T_xM$ can be decomposed with respect to that basis, $Y_j = y^\mu_j \partial_\mu$. A frame at $x$ is now a set of $n$ linearly independent vectors $Y_j \in T_xM$, $j=1,\dots,n$, parameterized by their coordinates $y^\mu_j$, where both $\mu$ and $j$ run from $1$ to $n$. Linear independence is equivalent to $\det y \neq 0$, and oriented frames have the same sign of $\det y$. The (oriented) frame bundle $F^+$ is now given by the base space $M$ with the set of smooth (positively oriented) frames attached to each point $x \in M$. A point in $F^+$ is thus (locally) given by the collection \[ (x,\{Y_j\})=(x^\mu, y_j^\mu)_{\mu,j=1,\dots,n}~, \qquad \det y > 0~, \] where $x^\mu$ are the coordinates of $x$ and the $y^\mu_j \in Gl^+(n)$ parameterize an oriented frame $\{Y_j\}_{j=1,\dots,n}$ at $x$. Here, $Gl^+(n)$ is the group of $n \times n$ matrices with positive determinant. In the overlap of two charts $U_1,U_2$, a point $x \in U_1 \cap U_2 \subset M$ will have coordinates $x^\mu$ in $U_1$ and $x^\nu{}'$ in $U_2$. The tangent vector $Y_i$ at a curve in $U_1 \cap U_2$ through $x$ is given \begin{eqnarray*} \mbox{in $U_1$ by}\quad{} & Y_i &= \frac{df(x^\mu(t))}{dt}\Big|_{t=0} = \frac{dx^\mu(t)}{dt} \Big|_{t=0} \; \partial_\mu f~, \\ \mbox{in $U_2$ by}\quad{} & Y_i &= \frac{df(x^\nu{}'(x^\mu(t)))}{dt}\Big|_{t=0} = \frac{dx^\mu(t)}{dt}\Big|_{t=0} \; \frac{\partial x^\nu{}'}{\partial x^\mu} \;\partial_\nu' f~, \end{eqnarray*} where $f$ is an arbitrary function on $F^+$. Hence, the coordinates $(x^\mu,y^\mu_j) \in U_1 \times Gl^+(n)$ and $(x^\nu{}',y^\nu_j{}') \in U_2 \times Gl^+(n)$ label the same point in $F^+$ iff $x^\mu,x^\nu{}'$ are the coordinates in $U_1,U_2$ of $x \in U_1 \cap U_2$ and $y^\nu_j{}' = (\partial x^\nu{}'/\partial x^\mu) y^\mu_j$. There is a natural action of $Gl^+(n)$ on a frame $\{Y_j\}$ at $x$: The matrix $g^i_j \in Gl^+(n)$ maps the vector $Y_i \in T_xM$ into the new vector $Y_i g^i_j=:Y'_j \in T_xM$, or~-- in coordinates~-- $y^\mu_i$ into $y^\mu_i g^i_j$. This $Gl^+(n)$-action extends naturally to the frame bundle, making $F^+$ to a $Gl^+(n)$-principal fibre bundle: \begin{equation} g:\,(x^\mu,y^\mu_i) \mapsto (x^\mu,y^\mu_i g^i_j)~. \label{right} \end{equation} The above action can be regarded as generated by a vector field according to the following construction. Let $gl(n)$ be the Lie algebra of $Gl^+(n)$. The exponential mapping assigns to $A \in gl(n)$ a curve $\exp(tA)$ in $Gl^+(n)$, which by (\ref{right}) induces a field of curves through every point of $F^+$. This field of curves provides us with a field of tangent vectors \[ \frac{df(x^\mu,y^\mu_i [\exp (tA)]^i_j}{dt} \Big|_{t=0} = \frac{\partial f}{\partial (y^\mu_k \delta^k_j)} \; \frac{d (y^\mu_i [\exp (tA)]^i_j)}{dt} \Big|_{t=0} = A^i_j y^\mu_i \frac{\partial}{\partial y^\mu_j} f ~, \] where $f$ is a function on $F^+$. Hence, each such vector field associated to $A \in gl(n)$ is generated by the following (vertical) vector fields \begin{equation} Y_i^j = y^\mu_i \frac{\partial}{\partial y^\mu_j} \equiv y^\mu_i \partial^j_\mu~. \label{Y} \end{equation} The vector field $A^\# = A^i_j Y^j_i$ associated to $A \in gl(n)$ is called the fundamental vector field corresponding to $A$. A somewhat tricky construction provides us with an ${\fam\bbfam R}^n$-valued $1$-form $\alpha$ on $F^+$, sometimes called soldering form or canonical $1$-form. A point $p =(x,\{y^\mu_j\}) \in F^+$ assigns to a vector $\tilde{V} \in T_xM$ a vector $\Phi_p(\tilde{V}) \in {\fam\bbfam R}^n$ by decomposing $\tilde{V}$ with respect to the basis $Y_j=y^\mu_j \partial_\mu$. Thus, $[\Phi_p(\tilde{V})]^j Y_j = \tilde{V}$. Now, the ${\fam\bbfam R}^n$-valued $1$-form $\alpha$ is defined by \begin{equation} \alpha(V) \big|_p = \Phi_p (\pi_* V)~, \qquad V \in T_p F^+~. \label{defalpha} \end{equation} By $\big|_p$ we denote the value of a differential form or a vector field at the point $p \in F^+$. In (\ref{defalpha}), $\pi_*$ is the differential of the vertical projection $\pi(x,\{y^\mu_j\}) = x$ which projects the vector $V=V^\mu \partial_\mu + V^\mu_j \partial^j_\mu \in T_p F^+$ into the vector $\pi_* V =V^\mu \partial_\mu \in T_{\pi(p)}M$. In this notation we have $\pi_* V = V^\nu (y^{-1})^j_\nu Y_j$, using the obvious definition $y^\mu_i(y^{-1})^i_\nu = \delta^\mu_\nu$. This gives $[\Phi_p (\pi_* V)]^j = V^\nu (y^{-1})^j_\nu$. On the other hand, decomposing $\alpha^i=\alpha^i_\mu dx^\mu + \alpha^{ik}_\mu dy^\mu_k$ and using the definition \[ dy^\mu_i(\partial^j_\nu) = \delta^j_i \delta^\mu_\nu~,\quad dx^\mu(\partial_\nu) = \delta^\mu_\nu~,\quad dy^\mu_i(\partial_\nu)=0~,\quad dx^\mu(\partial^j_\nu) =0 \] of the pairing between covectors and vectors, we have $\alpha^j(V) = \alpha^j_\mu V^\mu + \alpha^{ji}_\mu V^\mu_i$, giving \begin{equation} \alpha^j = (y^{-1})^j_\mu \,dx^\mu~. \label{alpha} \end{equation} The definition (\ref{defalpha}), although involving local coordinates in the construction, is independent of the choice of charts. Indeed, if $p \in F^+$ has the coordinates $(x^\mu,y^\mu_j)$ and $(x^\nu{}', y^\nu_j{}')$ in two charts $U_1 \times Gl^+(n)$ and $U_2 \times Gl^+(n)$ of $F^+$, with $y^\nu_j{}'=(\partial x^\nu{}'/\partial x^\mu) y^\mu_j$, then $\pi_* V = V^\mu \partial_\mu \in T_{\pi(p)} U_1$ and $\pi_* V = V^\nu{}' \partial_{\nu'} \in T_{\pi(p)} U_2$, with $ V^\nu{}' = (\partial x^\nu{}'/\partial x^\mu) V^\mu$. This means that $V^\mu (y^{-1})^j_\mu = V^\nu{}' (y^{-1})^j_{\nu'} \in {\fam\bbfam R}$ give the same value for $[\Phi_p (\pi_* V)]^j$. \subsection{Connection} A connection is the splitting of the tangent space $T_pF^+$ at $p \in F^+$ into a direct sum of a vertical space $V_pF^+$ (generated by $Y^i_j=y^\mu_j \partial^i_\mu$) and a horizontal space $H_pF^+$ such that $H_{pg}F^+=R_{g*} H_pF^+$. In the last equation, $pg \in F^+$ is the point obtained by the action (\ref{right}) of $g \in Gl^+(n)$ and $R_{g*}$ is the induced push-forward of a vector in $T_pF^+$ to a vector in $T_{pg}F^+$. If $V \in T_pF^+$ is the tangent vector of a curve $p(t)$ in $F^+$ through $p$, then the push-forward $R_{g*}V$ is the tangent vector of the curve $p(t)g$ through $pg$. Explicitly, let $f$ be a function on $F^+$ and $V=V_j^\mu \partial^j_\mu+ V^\mu \partial_\mu \in T_pF^+$ be tangent to the curve $C=(x^\mu(t),y^\mu_j(t))$ at $p$, i.e. \begin{equation} Vf=(df(p)/dt)=\big((dx^\mu/dt)\partial_\mu + (dy^\mu_j/dt)\partial^j_\mu\big)f~. \label{Vf} \end{equation} Then, the push-forward is given by \begin{equation} (R_{g*}V)f = \frac{df(p(t)\,g)}{dt} = \frac{dx^\mu(t)}{dt}\partial_\mu f + \frac{d (y^\mu_j(t)\, g^j_i)}{dt} \frac{\partial f}{\partial \hat{y}^\mu_i}~,\qquad \hat{y}^\mu_i := y^\mu_k g^k_i~. \label{RV} \end{equation} Thus, we obtain \begin{equation} R_{g*}V= V_j^\mu g^j_i \,\widehat{\partial^i_\mu} + V^\mu \partial_\mu~, \qquad \widehat{\partial^i_\mu} := \partial /\partial \hat{y}^\mu_i~. \label{RgV} \end{equation} In practice the connection is most conveniently characterized by the connection form $\omega$, a $gl(n)$-valued differential $1$-form with the following properties: For a given matrix $A \in gl(n)$ let $A^\#= A^i_j Y^j_i$ be the corresponding fundamental vector field. Then $\omega$ is defined by \begin{equation} \omega(A^\#) = A ~,\qquad \omega\big|_{pg}(R_{g*} V) = g^{-1} \Big(\omega\big|_p(V) \Big)g~, \label{omega} \end{equation} for $V \in T_pF^+$ and $g \in Gl^+(n)$. At the point $p=(x^\mu,y^\mu_i)$, the components $\omega^i_j$ of the connection form will have the decomposition \[ \omega^i_j = W^{ik}_{j \mu} dy^\mu_k + W^{i}_{j\mu} dx^\mu ~, \] for certain functions $W$. From (\ref{Y}) we get immediately \[ A^i_j \equiv \omega^i_j(A^\#) = W^{ik}_{j\mu} y^\mu_l A^l_k~, \] which gives $W^{ik}_{j\mu} = \delta^k_j (y^{-1})^i_\mu$. This suggests the following ansatz \begin{equation} \omega^i_j = (y^{-1})^i_\mu ( d y^\mu_j + \Gamma^\mu_{\nu\alpha} y^\nu_j d x^\alpha)~, \label{om} \end{equation} where $\Gamma^\mu_{\nu\alpha}$ is a so far undetermined function of $x$ and $y$. Using (\ref{RgV}) we write down \begin{eqnarray*} \omega^i_j\big|_{pg}(R_{g*}V) &=& (g^{-1})_k^i(y^{-1})^k_\mu \big( d \hat{y}^\mu_j + \Gamma^\mu_{\nu\alpha}\big|_{pg}\, y^\nu_k g^k_j d x^\alpha\big) \big( V_m^\nu g^m_l \widehat{\partial_\nu^l} + V^\mu \partial_\mu \big) \\ &=& (g^{-1})_k^i(y^{-1})^k_\mu V_m^\mu g^m_j + (g^{-1})_k^i(y^{-1})^k_\mu \Gamma^\mu_{\nu\alpha}\big|_{pg} \, V^\alpha y^\nu_k g^k_j ~. \end{eqnarray*} On the other hand, \begin{eqnarray*} \big[g^{-1}\big(\omega\big|_{p}(V)\big)g\big]^i_j = (g^{-1})_k^i(y^{-1})^k_\mu V_m^\mu g^m_j + (g^{-1})_k^i(y^{-1})^k_\mu \Gamma^\mu_{\nu\alpha}\big|_p \, V^\alpha y^\nu_k g^k_j ~. \end{eqnarray*} Thus, the condition (\ref{omega}) tells us that $\Gamma^\mu_{\nu\alpha}\big|_p=\Gamma^\mu_{\nu\alpha}\big|_{pg}$, {\em which means that $\Gamma^\mu_{\nu\alpha}$ depends only on the base point $x$.} Now, the horizontal vector fields $X_i$ associated to the connection are defined as the kernel of $\omega$ and the dual of $\alpha$, \begin{equation} \omega^i_j (X_k)=0~,\qquad \alpha^j(X_i)=\delta^j_i~. \end{equation} These equations are easy to solve: \begin{equation} X_i = y^\mu_i (\partial_\mu - \Gamma^\nu_{\alpha\mu} y^\alpha_j \partial^j_\nu)~. \label{X} \end{equation} The torsion form $\Theta$ on $F^+$ is an ${\fam\bbfam R}^n$-valued differential $2$-form defined as the covariant derivative of $\alpha$, \begin{equation} \Theta^i = d \alpha^i + \omega^i_j \wedge \alpha^j~. \end{equation} Using (\ref{alpha}) and (\ref{om}) we compute \begin{eqnarray*} \Theta^i &=& - (y^{-1})^i_\nu (y^{-1})^j_\mu \,d y^\nu_j \wedge dx^\mu + (y^{-1})^i_\mu ( d y^\mu_j + \Gamma^\mu_{\nu\alpha} y^\nu_j d x^\alpha) \wedge (y^{-1})^j_\beta dx^\beta \\ &=& (y^{-1})^i_\mu \Gamma^\mu_{\nu\alpha} d x^\alpha \wedge dx^\nu ~. \end{eqnarray*} The torsion vanishes iff the connection coefficients are symmetric, $\Gamma^\mu_{\nu\alpha} =\Gamma^\mu_{\alpha\nu}$. \subsection{Diffeomorphisms} Let $\psi$ be a local (orientation preserving) diffeomorphism of $M$. By push-forward it maps a frame $\{Y_j\}$ at $x \in M$ into the frame $\{\psi_* Y_j\}$ at $\psi(x) \in M$. If $Y$ is the tangent vector at $x$ of a curve $C=\{x^\mu(t)\}$ through $x$, then $\psi_* Y$ is the tangent vector at $\psi(x)$ of the curve $\psi(C)=\{\psi(x(t))^\nu\}$. We evaluate both vectors on a function $\phi$ on $M$: \begin{eqnarray*} Y \phi &=& \frac{d}{dt} \phi(x^\mu(t)) \Big|_{t=0} = \frac{\partial \phi(x^\mu)}{\partial x^\mu} \: \frac{dx^\mu(t)}{dt} \Big|_{t=0} = (dx^\mu(t)/dt)\big|_{t=0} \:\partial_\mu \phi~, \\ (\psi_* Y) \phi &=& \frac{d}{dt} \phi (\psi(x(t))^\nu) \Big|_{t=0} = \frac{\partial \phi(\tilde{x}^\nu)}{\partial \tilde{x}^\mu} \: \frac{\partial \psi(x)^\mu}{\partial x^\nu} \: \frac{dx^\nu(t)}{dt} \Big|_{t=0} \\ &=& \partial_\nu \psi(x)^\mu\: (dx^\nu(t)/dt)\big|_{t=0} \: \widetilde{\partial_\mu} \phi~, \end{eqnarray*} with $\tilde{x}^\mu = \psi(x)^\mu$ and $\widetilde{\partial_\mu} =\partial/\partial \tilde{x}^\mu$. Recall that $\partial_\mu$ and $\widetilde{\partial_\mu}$ are the bases of vector fields in $T_xM$ and $T_{\psi(x)}M$, respectively. Hence, if $y^\mu_j$ are the coordinates of $Y_j \in T_xM$, then $\psi_* Y_j \in T_{\psi(x)}M$ has the coordinates $\partial_\nu \psi(x)^\mu \, y^\nu_j$, with respect to these bases. To summarize, the action $\tilde{\psi}$ on $F^+$ of a diffeomorphism $\psi$ of $M$ is given by \begin{equation} \tilde{\psi}: (\{x^\mu\}, \{y^\mu_i\}) \mapsto (\{\tilde{x}^\mu := \psi(x)^\mu\}, \{\tilde{y}^\mu_i := \partial_\nu \psi(x)^\mu \, y^\nu_i\}) ~. \label{psi} \end{equation} Note that the (right) action (\ref{right}) of $Gl^+(n)$ on $F^+$ and the (left) action (\ref{psi}) on $F^+$ of a diffeomorphism of $M$ commute with each other. We consider now the effect of a diffeomorphism $\psi$ of $M$ on the horizontal vector fields $X_i$. We use the following \begin{Lemma} The pull-back $\tilde{\psi}^* \omega$ of the connection form via the action (\ref{psi}) of the induced diffeomorphism $\tilde{\psi}$ of $F^+$ is again a connection form. \end{Lemma} {\it Proof.} We start from (\ref{Vf}) and compute \begin{eqnarray} (\tilde{\psi}_* V) f &=& \frac{d}{dt} f\Big(\tilde{x}^\mu(x^\nu(t)), \tilde{y}^\mu_i(y^\nu_j(t),x^\nu(t))\Big) \nonumber \\ &=& \frac{\partial f}{\partial \tilde{x}^\mu} \frac{\partial \psi(x)^\mu}{\partial x^\alpha} \frac{dx^\alpha}{dt} + \frac{\partial f}{\partial \tilde{y}^\mu_i} \Big( \frac{\partial^2 \psi(x)^\mu}{\partial x^\alpha \partial x^\beta} y_i^\beta \frac{dx^\alpha}{dt} + \partial_\alpha \psi(x)^\mu \frac{d y^\alpha_i}{dt} \Big) \nonumber \\ &=& \partial_\alpha \psi(x)^\mu V^\alpha \widetilde{\partial_\mu} f + \big( \partial_\alpha \partial_\beta \psi(x)^\mu y_i^\beta V^\alpha + \partial_\alpha \psi(x)^\mu V^\alpha_i \big) \widetilde{\partial^i_\mu} f~, \label{psiV} \end{eqnarray} where $\widetilde{\partial^i_\mu}= \frac{\partial}{\partial \tilde{y}^\mu_i}$. For $V=A^\#$ we have $V^\mu=0$ and $V^\mu_i=A^j_i y^\mu_j$, see (\ref{Y}). This means \begin{equation} \big(\tilde{\psi}_* A^\# \big)\big|_{\tilde{\psi}(p)}= A^j_i \partial_\alpha \psi(x)^\mu y^\alpha_j \widetilde{\partial^i_\mu} = A^j_i \widetilde{y^\alpha_j} \widetilde{\partial^i_\mu} = A^\# \big|_{\tilde{\psi}(p)}~. \label{Ainv} \end{equation} {\it The fundamental vector field $A^\#$ is invariant under diffeomorphisms.} This gives \[ (\tilde{\psi}^*\omega)(A^\#)\big|_p=\omega(\tilde{\psi}_*A^\#) \big|_{\tilde{\psi}(p)} =\omega(A^\#) \big|_{\tilde{\psi}(p)} = A~. \] The second identity to prove is \begin{eqnarray*} &&(\tilde{\psi}^* \omega)\big|_{pg}(R_{g*}V)\big|_{pg} = g^{-1} \Big((\tilde{\psi}^* \omega)(V)\big|_p \Big) g \quad \\ && \Rightarrow \quad \omega \big|_{\tilde{\psi}(pg)} \big( \psi_* \big((R_{g*}V)\big|_{pg} \big) \big|_{\tilde{\psi}(pg)} \big) = g^{-1} \Big( \omega\big|_{\tilde{\psi}(p)} (\tilde{\psi}_* V) \big|_{\tilde{\psi}(p)} \Big) g ~. \end{eqnarray*} According to (\ref{RV}) we replace $V^\mu_i$ by $V^\mu_j g^j_i$ and $y^\mu_i$ by $y^\mu_j g^j_i$ and insert this into (\ref{psiV}): \[ \tilde{\psi}_* \big(R_{g*}V \big)\big|_{\tilde{\psi}(pg)} = \partial_\alpha \psi(x)^\mu V^\alpha \widetilde{\partial_\mu} + \big( \partial_\alpha \partial_\beta \psi(x)^\mu y_j^\beta V^\alpha + \partial_\alpha \psi(x)^\mu V^\alpha_j \big) g^j_i \widetilde{\widehat{\partial^i_\mu}} ~, \] where $\widetilde{\widehat{\partial^i_\mu}}= \partial/\partial \widetilde{\widehat{y^\mu_i}}$ and $\widetilde{\widehat{y^\mu_i}}= \partial_\alpha \psi(x)^\mu y^\alpha_j g^j_i$. We must now evaluate \[ \omega^a_b \big|_{\tilde{\psi}(pg)} = (g^{-1})^a_c (y^{-1})^c_\gamma ((\partial\psi(x))^{-1})^\gamma_\delta \Big( d \widetilde{\widehat{y^\delta_b}} + \Gamma^\delta_{\epsilon\zeta}\big|_{\psi(x)} \, \partial_\eta\psi(x)^\epsilon y^\eta_d g^d_b \, d \widetilde{x^\zeta} \Big) \] on the above vector: \begin{eqnarray} && \omega^a_b \big|_{\tilde{\psi}(pg)} \big(\tilde{\psi}_* \big(R_{g*}V \big)\big|_{\tilde{\psi}(pg)} \big) \nonumber \\ &&= (g^{-1})^a_c \Big\{(y^{-1})^c_\gamma ((\partial\psi(x))^{-1})^\gamma_\delta \Big( \big( \partial_\alpha \partial_\beta \psi(x)^\delta y_d^\beta V^\alpha + \partial_\alpha \psi(x)^\delta V^\alpha_d \big) \nonumber \\ &&{} \hspace*{5cm} + \Gamma^\delta_{\epsilon\zeta}\big|_{\psi(x)} \, \partial_\eta\psi(x)^\epsilon y^\eta_d \, \partial_\nu \psi(x)^\zeta V^\nu \Big) \Big\} g^d_b ~. \label{qed} \end{eqnarray} Taking $g=e$ (identity matrix), it is obvious that the term in braces $\{~\}$ equals $\omega^c_d (\tilde{\psi}_* V) \big|_{\tilde{\psi}(p)}$, which finishes the proof of the Lemma. \hfill \qed \vskip 2ex We can now rewrite the term in braces in (\ref{qed}) in a slightly different way: {\arraycolsep 0pt \begin{eqnarray*} && \omega^c_d (\tilde{\psi}_* V)\big|_{\tilde{\psi}(p)} = \big(\tilde{\psi}^* \omega^c_d\big) (V) \big|_p \\ &&= (y^{-1})^c_\gamma V^\gamma_d + (y^{-1})^c_\gamma ((\partial\psi(x))^{-1})^\gamma_\delta \Big( \partial_\alpha \partial_\beta \psi(x)^\delta + \Gamma^\delta_{\epsilon\zeta}\big|_{\psi(x)} \, \partial_\beta\psi(x)^\epsilon \partial_\alpha \psi(x)^\zeta \Big) y_d^\beta V^\alpha \\ &&= (y^{-1})^c_\gamma ( d y^\gamma_d + \tilde{\Gamma}^\gamma_{\beta\alpha}\big|_x\, y^\beta_d d x^\alpha)(V^\nu_k \partial^k_\nu + V^\nu \partial_\nu)~, \end{eqnarray*}}% where $\tilde{\Gamma}^\gamma_{\beta\alpha}$ are the connection coefficients of the connection $\tilde{\psi}^* \omega$. This provides us with the following transformation law for the connection coefficients: \begin{equation} \tilde{\Gamma}^\gamma_{\beta\alpha}\big|_x = ((\partial\psi(x))^{-1})^\gamma_\delta \, \Gamma^\delta_{\epsilon\zeta}\big|_{\psi(x)} \, \partial_\beta\psi(x)^\epsilon \partial_\alpha \psi(x)^\zeta + ((\partial\psi(x))^{-1})^\gamma_\delta \, \partial_\alpha \partial_\beta \psi(x)^\delta ~. \label{Gamma} \end{equation} Now there is an immediate question to ask: Which are the horizontal vector fields $\tilde{X}_i$ to the new connection form $\tilde{\psi}^* \omega$? We have \[ 0 = (\tilde{\psi}^* \omega)\big|_p (\tilde{X}_i \big|_p) = \omega \big|_{\tilde{\psi}(p)} (\tilde{\psi}_* \tilde{X}_i) \big|_{\psi(p)} ~, \] which tells us \begin{equation} \tilde{X}_i \big|_p = \tilde{\psi}^{-1}_* \big(X_i\big|_{\tilde{\psi}(p)} \big) = y^\mu_i (\partial_\mu - \tilde{\Gamma}^\nu_{\alpha\mu} \big|_x\,y^\alpha_j \partial^j_\nu)~. \label{tX} \end{equation} The action (\ref{psi}) preserves the ${\fam\bbfam R}^n$-valued $1$-form $\alpha$ given in (\ref{defalpha}) and (\ref{alpha}). Indeed, for $V = V^\mu \partial_\mu + V^\mu_i \partial^i_\mu \in T_pF^+$ we compute using (\ref{psiV}) \begin{eqnarray*} (\tilde{\psi}^* \alpha^j)\big|_p (V) &=& \alpha\big|_{\tilde{\psi}(p)} (\tilde{\psi}_*V) \\ &=& (\tilde{y}^{-1})^j_\mu d \tilde{x}^\mu \Big( \partial_\alpha \psi(x)^\mu V^\alpha \widetilde{\partial_\mu} + \big( \partial_\alpha \partial_\beta \psi(x)^\mu y_i^\beta V^\alpha + \partial_\alpha \psi(x)^\mu V^\alpha_i \big) \widetilde{\partial^i_\mu} \Big) \\ &=& (\tilde{y}^{-1})^j_\mu \partial_\alpha \psi(x)^\mu V^\alpha = (y^{-1})^j_\alpha V^\alpha \equiv \alpha^j\big|_p (V)~. \end{eqnarray*} \section{Crossed product} \label{cross} The properties listed in Proposition \ref{prpgen} and derived throughout section \ref{general} are the basis for the construction of the Hopf algebra of Connes and Moscovici \cite{cm2}. The idea is to apply the vertical and horizontal vector fields $Y^j_i$ and $X_i$ to a crossed product ${\cal A}$ defined below and to derive their coproduct from \begin{equation} X_i (ab) = \Delta(X_i)\,(a \otimes b)~,\qquad Y^j_i (ab) = \Delta(Y^j_i)\,(a \otimes b)~,\qquad a,b \in {\cal A}~. \label{gendelta} \end{equation} We refer to \cite{m} for an introduction to Hopf algebras and related topics. Let $\Gamma$ be the pseudogroup of local (orientation preserving) diffeomorphisms of $M$. We consider the crossed product of the algebra $C^\infty_c(F^+)$ of smooth functions with compact support on the frame bundle $F^+$ by the action of $\Gamma$, \begin{equation} {\cal A} = C^\infty_c(F^+) \mathop{>\!\!\!\triangleleft} \Gamma~. \end{equation} As a set, ${\cal A}$ can be regarded as the tensor product of $C^\infty_c(F^+)$ with $\Gamma$. It is generated by the monomials \begin{equation} f U^*_\psi~,\qquad f \in C^\infty_c({\rm Dom}(\tilde{\psi}))~,\quad \psi \in \Gamma~, \end{equation} where $\tilde{\psi}$ is the diffeomorphism of $F^+$ induced by $\psi \in \Gamma$ according to (\ref{psi}). As an algebra, the multiplication rule in ${\cal A}$ is defined by \begin{equation} f_1 U^*_{\psi_1} \; f_2 U^*_{\psi_2} := f_1 (f_2 \circ \tilde{\psi}_1) U^*_{\psi_2 \psi_1} ~. \label{prod} \end{equation} In this formula, the function $f_1 (f_2 \circ \tilde{\psi}_1) \in C^\infty_c(D_{\psi_1,\psi_2})$, with $D_{\psi_1,\psi_2}:= {\rm Dom}(\tilde{\psi}_1) \cap \tilde{\psi}_1^{-1}({\rm Dom} (\tilde{\psi}_2)) \subset F^+$, maps $p \in D_{\psi_1,\psi_2}$ into $f_1(p) \,f_2 (\tilde{\psi}_1(p)) \in {\fam\bbfam R}$ (or ${\fam\bbfam C})$. The star on $U_\psi^*$ refers to the contravariant multiplication rule $ U^*_{\psi_1} U^*_{\psi_2} = U^*_{\psi_2 \psi_1}$. Associativity of ${\cal A}$ follows -- for appropriate support of the functions -- from \[ \big(f_1 (f_2 \circ \psi_1)\big) (f_3 \circ (\psi_2\psi_1)) = f_1 \big((f_2 (f_3 \circ \psi_2)) \circ \psi_1 \big)~. \] We consider now the action of the vertical and horizontal vector fields $Y^j_i$ and $X_i$ described in section \ref{general} on the algebra ${\cal A}$. That action is simply defined as the action of the vector fields on the functions, \begin{equation} Y^j_i ( f U_\psi^*) = Y^j_i (f) U_\psi^*~,\qquad X_i ( f U_\psi^*) = X_i (f) U_\psi^*~. \end{equation} The interesting effects we are looking for are obtained by application of these vector fields to the product (\ref{prod}). For any vector field $V$ on $F^+$ we compute \begin{eqnarray} V(f_1 U^*_{\psi_1} \; f_2 U^*_{\psi_2}) \big|_p &=& V(f_1 (f_2 \circ \tilde{\psi}_1)) U^*_{\psi_2 \psi_1} \big|_p \nonumber \\ &=& \{V(f_1) \big|_p\; (f_2 \circ \tilde{\psi}_1) \big|_p + f_1 \big|_p \; V (f_2 \circ \tilde{\psi}_1) \big|_p\} U^*_{\psi_2 \psi_1}\nonumber \\ &=& V(f_1) U^*_{\psi_1} \big|_p\; f_2 U^*_{\psi_2} + f_1\big|_p \; ((\tilde{\psi}_{1*} V) f_2 ) \big|_{\tilde{\psi}_1(p)} U^*_{\psi_2 \psi_1} \nonumber \\ &=& V(f_1) U^*_{\psi_1} \big|_p\; f_2 U^*_{\psi_2} + f_1\big|_p \; U^*_{\psi_1} U^*_{\psi_1^{-1}} ((\tilde{\psi}_{1*} V) f_2) \big|_{\tilde{\psi}_1(p)} U^*_{\psi_2 \psi_1} \nonumber \\ &=& V(f_1) U^*_{\psi_1} \big|_p\; f_2 U^*_{\psi_2} + f_1 U^*_{\psi_1} \; ((\tilde{\psi}_{1*} V) f_2 )\big|_{\tilde{\psi}_1^{-1} \circ \tilde{\psi}_1(p)} U^*_{\psi_2} \nonumber \\ &=& V(f_1 U^*_{\psi_1}) \big|_p\; f_2 U^*_{\psi_2} + f_1 U^*_{\psi_1} \; \big(\tilde{\psi}_{1*} (V\big|_{\tilde{\psi}_1^{-1}(p)}) \big) f_2 U^*_{\psi_2} \big|_p~. \label{Vab} \end{eqnarray} In the third line we have used the definition of the push-forward. In the fifth line we have commuted $U^*_{\psi_1^{-1}}$ with the function $(\tilde{\psi}_{1*} V) f_2$, evaluated at $\tilde{\psi}_1(p)$. According to (\ref{prod}), after taking $U^*_{\psi_1^{-1}}$ to the right we must evaluate the function $(\tilde{\psi}_{1*} V) f_2$ at $\tilde{\psi}_1^{-1}(\tilde{\psi}_1(p))=p$. This means that the original field $V$ to push forward must be taken at $\tilde{\psi}_1^{-1}(p)$. Taking for $V$ the vertical vector fields $Y^j_i$ and recalling their invariance under diffeomorphisms (\ref{Ainv}), we obtain immediately \begin{equation} Y^j_i(ab) = Y^j_i(a) \,b + a\, Y^j_i(b) ~,\qquad a,b \in {\cal A}~. \label{Yab} \end{equation} The behavior of the horizontal vector fields $X_i$ is very different, because they do not commute with the diffeomorphisms. Eq.\ (\ref{tX}) tells us that if $X_i$ is horizontal to $\omega$, then $X_i^{(\psi_1)}:=\tilde{\psi}_{1*}(X_i\big|_{\tilde{\psi}_1^{-1}(p)})$ is horizontal to $(\tilde{\psi}^{-1}_1)^*\omega$. We denote the connection coefficients of $(\tilde{\psi}^{-1}_1)^*\omega$ by $\hat{\Gamma}^\nu_{\alpha\mu}$. We observe from (\ref{X}) and (\ref{Y}) that \begin{eqnarray} (X_i^{(\psi_1)} -X_i )\big|_p = (\Gamma^\nu_{\alpha\mu}\big|_x -\hat{\Gamma}^\nu_{\alpha\mu}\big|_x) y^\mu_i y^\alpha_j \partial^j_\nu &=& (\Gamma^\nu_{\alpha\mu}\big|_x -\hat{\Gamma}^\nu_{\alpha\mu}\big|_x) y^\mu_i y^\alpha_j (y^{-1})^k_\nu Y^j_k|_p \nonumber \\ &=:& \hat{\gamma}^k_{ji}\big|^{(\psi_1)}_p Y^j_k\big|_p~. \end{eqnarray} This gives from (\ref{Vab}) for the horizontal fields $X_i$ \begin{eqnarray} X_i(f_1 U^*_{\psi_1} \; f_2 U^*_{\psi_2}) \big|_p &=& X_i(f_1 U^*_{\psi_1}) \big|_p \; f_2 U^*_{\psi_2} \big|_p + f_1 U^*_{\psi_1}\big|_p \; X_i^{(\psi_1)} (f_2 U^*_{\psi_2})\big|_p \nonumber \\ &=& X_i(f_1 U^*_{\psi_1}) \big|_p \; f_2 U^*_{\psi_2} \big|_p + f_1 U^*_{\psi_1} \big|_p \; X_i (f_2 U^*_{\psi_2}) \big|_p \nonumber \\ && + f_1 U^*_{\psi_1} \big|_p \; \hat{\gamma}^k_{ij} \big|^{(\psi_1)}_p Y^j_k (f_2 U^*_{\psi_2}) \big|_p \nonumber \\ &=& X_i(f_1 U^*_{\psi_1}) \big|_p \; f_2 U^*_{\psi_2} \big|_p + f_1 U^*_{\psi_1} \big|_p \; X_i (f_2 U^*_{\psi_2}) \big|_p \nonumber \\ && + f_1 \big|_p \hat{\gamma}^k_{ij} \big|^{(\psi_1)}_{\tilde{\psi}_1(p)} U^*_{\psi_1} \; Y^j_k (f_2 U^*_{\psi_2}) \big|_p ~. \label{Xab} \end{eqnarray} Our goal is to express $\hat{\gamma}^k_{ij} \big|^{(\psi_1)}_{\tilde{\psi}_1(p)}$ in terms of some function evaluated at $p$. From (\ref{alpha}) and (\ref{om}) we conclude \begin{equation} \omega^k_j \big|_p - ((\tilde{\psi}^{-1})^*\omega^k_j) \big|_p = \hat{\gamma}^k_{ji} \big|^{(\psi)}_p\; \alpha^i\big|_p~. \label{hatgamma} \end{equation} We take this identity at $\tilde{\psi}(p)$ and apply $\tilde{\psi}^*$, which gives \begin{equation} (\tilde{\psi}^* \omega^k_j) \big|_p - \omega^k_j \big|_p = \hat{\gamma}^k_{ji} \big|^{(\psi)}_{\tilde{\psi}(p)}\; (\tilde{\psi}^* \alpha^i) \big|_p = \hat{\gamma}^k_{ji} \big|^{(\psi)}_{\tilde{\psi}(p)}\; \alpha^i \big|_p~, \label{hgp} \end{equation} using the invariance of $\alpha^i$ under diffeomorphisms in the last step. Replacing in (\ref{hgp}) $\psi$ by $\psi^{-1}$ and comparing with (\ref{hatgamma}) we get \begin{equation} \hat{\gamma}^k_{ji} \big|^{(\psi)}_{\tilde{\psi}(p)} = - \hat{\gamma}^k_{ji} \big|^{(\psi^{-1})}_p =: \gamma^k_{ji} \big|^{(\psi)}_p = (\tilde{\Gamma}^\nu_{\alpha\mu}\big|_x - \Gamma^\nu_{\alpha\mu}\big|_x) y^\alpha_j y^\mu_i (y^{-1})^k_\nu~, \label{gamma} \end{equation} where $\tilde{\Gamma}^\nu_{\alpha\mu}$ and $\Gamma^\nu_{\alpha\mu}$ are the connection coefficients of the connections $\tilde{\psi}^*\omega$ and $\omega$, respectively. Since $\tilde{\Gamma}^\nu_{\alpha\mu}$ is defined by the diffeomorphism $\psi$, we define an operator $\delta^k_{ji}$ on ${\cal A}$ by \begin{equation} \delta^k_{ji} (f U^*_\psi)\big|_{p} = \gamma^k_{ji}\big|^{(\psi)}_p\; f U^*_\psi\big|_{p} \label{deltaf} \end{equation} and get from (\ref{Xab}) and (\ref{gamma}) \begin{equation} X_i(ab) = X_i(a)\,b + a\,X_i(b) + \delta^k_{ji}(a)\,Y^j_k(b)~,\qquad a,b \in {\cal A}~. \label{Xiab} \end{equation} Next, we compute \begin{equation} \delta^k_{ji} (f_1 U^*_{\psi_1} \; f_2 U^*_{\psi_2}) \big|_p = \delta^k_{ji}(f_1 (f_2 \circ \tilde{\psi}_1) U^*_{\psi_2\psi_1}) \big|_p = \gamma^k_{ji}\big|^{(\psi_2\psi_1)}_p \, f_1\big|_p \,f_2\big|_{\tilde{\psi}_1(p)} U^*_{\psi_2\psi_1}~. \label{deltaab} \end{equation} Starting with (\ref{hgp}) and (\ref{gamma}) we compute \begin{eqnarray*} \gamma^k_{ji}\big|^{(\psi_2\psi_1)}_p \;\alpha^i \big|_p &=& (\tilde{\psi}_2\tilde{\psi}_1)^*(\omega^k_j \big|_{(\tilde{\psi}_2\tilde{\psi}_1)(p)}) - \omega^k_j\big|_p \\ &=& \tilde{\psi}_1^* \Big( \tilde{\psi}_2^* (\omega^k_j \big|_{(\tilde{\psi}_2\tilde{\psi}_1)(p)}) -\omega^k_j \big|_{\tilde{\psi}_1(p)} \Big) + \Big( \tilde{\psi}_1^* (\omega^k_j\big|_{\tilde{\psi}_1(p)} ) - \omega^k_j \big|_p \Big) \\ &=& \psi_1^* \Big(\gamma^k_{ji}\big|^{(\psi_2)}_{\tilde{\psi}_1(p)} \alpha^i \big|_{\tilde{\psi}_1(p)} \Big) + \gamma^k_{ji} \big|^{(\psi_1)}_p \alpha^i\big|_p \\ &=& \Big( \gamma^k_{ji}\big|^{(\psi_1)}_p + \gamma^k_{ji} \big|^{(\psi_2)}_{\tilde{\psi_1}(p)} \Big) \alpha^i\big|_p~. \end{eqnarray*} We used again the invariance of $\alpha^i$ under diffeomorphisms in the last line. We insert this result into (\ref{deltaab}) and get \begin{eqnarray*} \delta^k_{ji} (f_1 U^*_{\psi_1} \; f_2 U^*_{\psi_2}) \big|_p &=& \gamma^k_{ji}\big|^{(\psi_1)}_p f_1\big|_p \, f_2\big|_{\tilde{\psi}_1(p)} U^*_{\psi_2 \psi_1} + f_1\big|_p \, \gamma^k_{ji}\big|^{(\psi_2)}_{\tilde{\psi}_1(p)}\, f_2\big|_{\tilde{\psi}_1(p)} U^*_{\psi_2 \psi_1} \\ &=& \gamma^k_{ji}\big|^{(\psi_1)}_p f_1\big|_p U^*_{\psi_1}\; f_2\big|_p U^*_{\psi_2} + f_1\big|_p U^*_{\psi_1} \;\gamma^k_{ji}\big|^{(\psi_2)}_p f_2\big|_p U^*_{\psi_2}~, \end{eqnarray*} which means \begin{equation} \delta^k_{ji}(ab) = \delta^k_{ji}(a)\,b + a\,\delta^k_{ji}(b)~. \label{delta} \end{equation} The equations (\ref{Yab}), (\ref{Xiab}) and (\ref{delta}) endow the operators $X_i,Y_k^j$ and $\delta^k_{ji}$ with the structure of a coalgebra, with the coproduct (\ref{gendelta}) given by \begin{eqnarray} \Delta(Y^j_k) &=& Y^k_j \otimes 1 + 1 \otimes Y^j_k~, \nonumber \\ \Delta(X_i) &=& X_i \otimes 1 + 1 \otimes X_i + \delta^k_{ji} \otimes Y^j_k~, \label{Delta} \\ \Delta(\delta^k_{ji}) &=& \delta^k_{ji} \otimes 1 + 1 \otimes \delta^k_{ji}~, \nonumber \\ \Delta(1) &=& 1 \otimes 1~, \nonumber \end{eqnarray} with $1$ being the identity on ${\cal A}$. It is easy to check that $\Delta$ is coassociative on the linear space ${\fam\bbfam R}(1,X_i,Y_k^j,\delta^k_{ji})$, \begin{equation} (\Delta \otimes {\rm id}) \circ \Delta = ({\rm id} \otimes \Delta) \circ \Delta~. \label{coass} \end{equation} \section{From Lie algebra to Hopf algebra} \label{Lie} Vector fields form a Lie algebra, so it is natural to investigate whether $X_i,Y_k^j,\delta^k_{ji}$ generate a Lie algebra. We compute the mutual commutators, starting with $Y^i_j$: \begin{eqnarray} [Y^i_j,Y^k_l] (fU^*_\psi) &=& (y^\mu_j \partial^i_\mu y^\nu_l \partial^k_\nu - y^\nu_l \partial^k_\nu y^\mu_j \partial^i_\mu ) fU^*_\psi \nonumber \\ &=& (\delta^i_l Y^k_j - \delta^k_j Y^i_l) (fU^*_\psi)~, \\{} [Y^k_j,X_i] (fU^*_\psi) &=& \big( y^\mu_j \partial^k_\mu (y^\nu_i \partial_\nu - \Gamma^\beta_{\alpha\nu} y^\nu_i y^\alpha_l \partial^l_\beta) -(y^\nu_i \partial_\nu - \Gamma^\beta_{\alpha\nu} y^\nu_i y^\alpha_l \partial^l_\beta) y^\mu_j \partial^k_\mu\big) fU^*_\psi \nonumber \\ &=& \delta^k_i X_j (fU^*_\psi)~, \\{} [Y^i_j,\delta^k_{lm}] (fU^*_\psi) &=& \big( y^\mu_j \partial^i_\mu (\tilde{\Gamma}^\nu_{\beta\alpha}{-} \Gamma^\nu_{\beta\alpha}) y^\beta_l y^\alpha_m (y^{-1})^k_\nu - (\tilde{\Gamma}^\nu_{\beta\alpha}{-}\Gamma^\nu_{\beta\alpha}) y^\beta_l y^\alpha_m (y^{-1})^k_\nu y^\mu_j \partial^i_\mu \big) fU^*_\psi \nonumber \\ &=& (\delta^i_l \delta^k_{jm} + \delta^i_m \delta^k_{lj} - \delta^k_j \delta^i_{lm}) ( fU^*_\psi)~. \label{Ydel} \end{eqnarray} So far we have considered the most general connection on $M$, even with torsion. But now, the commutator of horizontal vector fields \begin{eqnarray} [X_i,X_j] &=& R^k_{lij} Y^l_k + \Theta^k_{ij} X_k~, \\ R^k_{lij} &=& (y^{-1})^k_\sigma y^\rho_l y^\mu_i y^\nu_j \big( \partial_\nu\Gamma^\sigma_{\rho\mu}- \partial_\mu\Gamma^\sigma_{\rho\nu} + \Gamma^\beta_{\rho\mu} \Gamma^\sigma_{\beta\nu} - \Gamma^\beta_{\rho\nu} \Gamma^\sigma_{\beta\mu} \big)~, \nonumber \\ \Theta^k_{ij} &=& (y^{-1})^k_\rho y^\mu_i y^\nu_j \big( \Gamma^\rho_{\mu\nu} - \Gamma^\rho_{\nu\mu}\big)~, \nonumber \end{eqnarray} leads to curvature $R$ and torsion $\Theta$, i.e.\ not to structure `constants'. Torsion can be avoided by the choice of the connection, but we would be forced to include $R^k_{lij} Y^l_k$ and its repeated commutators with $X_m$ in the list of generators of the Lie algebra we are looking for. To avoid these terms we follow \cite{cm2} and restrict ourselves to a {\it flat} manifold. Locally this is always possible, and globally it is achieved via the {\it Morita equivalence}. For a locally finite cover of the manifold $M$ by charts $U_\alpha$, let $N=\coprod U_\alpha$ be the disjoint union of the charts. Moreover, let $\Gamma'$ be the pseudogroup of local diffeomorphisms of $N$. Without giving the proof we recall from \cite{cm2} that the two algebras ${\cal A} = C^\infty_c(F^+(M)) \mathop{>\!\!\!\triangleleft} \Gamma$ and ${\cal A}' = C^\infty_c(F^+(N)) \mathop{>\!\!\!\triangleleft} \Gamma'$ are Morita equivalent. There is a canonical connection on $N$, the flat connection given by $\Gamma^\alpha_{\beta\gamma}=0$. This means that given $M$ we pass to $N$ and the corresponding crossed product ${\cal A}'$ and derive there the coproduct and Lie algebra structure of vector fields on $F^+(N)$ for the flat connection. Thus, the horizontal vector fields take the simple form $X_i = y^\mu_i \partial_\mu$, and they now commute with each other: \begin{equation} [X_i,X_j] (f U^*_\psi) = 0 ~. \end{equation} Due to (\ref{Gamma}), (\ref{gamma}) and (\ref{deltaf}), the action of $\delta^k_{ji}$ on ${\cal A}$ simplifies in the case of a flat manifold to \begin{equation} \delta^k_{ji} (f U^*_\psi) = ((\partial\psi(x))^{-1})^\nu_\beta\, \partial_\mu\partial_\alpha \psi(x)^\beta \, y^\mu_j y^\alpha_i (y^{-1})^k_\nu\, f U^*_\psi~. \end{equation} The (repeated) commutator with $X_l$ leads to new operators on ${\cal A}$, \begin{eqnarray} \delta^k_{ji,l_1\dots l_n} (f U^*_\psi) &:=& [X_{l_n},\dots ,[X_{l_1},\delta^k_{ji}]\dots] (f U^*_\psi) \label{flat} \\ \nonumber &=& \partial_{\lambda_n} \dots \partial_{\lambda_1} \Big(((\partial\psi(x))^{-1})^\nu_\beta\, \partial_\mu\partial_\alpha \psi(x)^\beta \Big) y^\mu_j y^\alpha_i (y^{-1})^k_\nu\, y^{\lambda_1}_{l_1} \cdots y^{\lambda_n}_{l_n} \; f U^*_\psi~. \end{eqnarray} It is clear that all these operators $\delta$ commute with each other, \begin{equation} [\delta^k_{ji,l_1\dots l_n},\delta^c_{ba,d_1\dots d_n}](f U^*_\psi) =0~. \end{equation} We see that the linear space generated by $X_i,Y^k_j, \delta^k_{ji,l_1\dots l_n}$ forms a Lie algebra, and we let ${\cal H}$ be the corresponding enveloping algebra. This is the algebra of polynomials in the generators of the Lie algebra, with the commutation relations inherited from the Lie algebra. Thus a (Poincar\'e-Birkhoff-Witt) basis in ${\cal H}$ is given by \[ X_{i_1} \cdots X_{i_\alpha} Y^{k_1}_{j_1} \cdots Y^{k_\beta}_{j_\beta} \delta^{a_1}_{b_1 c_1} \cdots \delta^{a_\gamma}_{b_\gamma c_\gamma} \delta^{d_1}_{e_1 f_1,h_1} \cdots \delta^{d_\delta}_{e_\delta f_\delta,h_\delta} \cdots ~, \] with $i_1\leq i_2\leq \dots \leq i_n$ and so on for the other indices. We extend the coproduct (\ref{Delta}) recursively to ${\cal H}$ by the definition \begin{equation} \Delta (h^1 h^2) = \Delta (h^1) \, \Delta(h^2) := \sum h_1^1 h^1_2 \otimes h_1^2 h^2_2 ~,\qquad \Delta(h_i) = \sum h^1_i \otimes h^2_i~, \label{copro} \end{equation} for $h_1,h_2 \in {\cal H}$. The coproduct is automatically coassociative (\ref{coass}) and by construction (\ref{copro}) compatible with the multiplication in ${\cal H}$. For notational convenience we abbreviate $\delta^A=\delta^k_{ji}$ with $A=1,\dots, n^2(n+1)/2$, due to symmetry in $i,j$. Moreover, we introduce a string $a=a_1a_2\dots a_k$ for the repeated commutators with $X_{a_1},\dots,X_{a_k}$ and denote its length by $|a|=k$. Next, let ${\cal H}_n$ be the commutative algebra of polynomials in the variables $1$ and $\delta^A_a$, with $0\leq |a| \leq n$. Let ${\cal H}^0_n$ be the ideal of polynomials vanishing at $0$. We obtain a more explicit formula of the coproduct in \begin{Lemma} $\Delta \delta^A_a = \delta^A_a \otimes 1 + 1 \otimes \delta^A_a + R^A_a~, \qquad R^A_a \in {\cal H}^0_{n-1} \otimes {\cal H}^0_{n-1} \quad \mbox{for} ~~ |a|=n~.$ \label{dex} \end{Lemma} {\it Proof.} The Lemma holds for $n=0$ with $R^A=0$. Assuming it holds for $|a|=n$ we compute for $b:=ai$ (appending the index $i$ to the string $a$), $|b|=n+1$, \begin{eqnarray} \Delta(\delta^A_b) &=& \Delta([X_i,\delta^A_a]) = [ \Delta(X_i),\Delta(\delta^A_a)] \nonumber \\ &=& [X_i \otimes 1 + 1 \otimes X_i + \delta^k_{ji} \otimes Y^j_k , \delta^A_a \otimes 1 + 1 \otimes \delta^A_a + R^A_a] \nonumber \\ &=& \delta^A_b \otimes 1 + 1 \otimes \delta^A_b + R^A_b~, \quad\mbox{with} \nonumber \\[1ex] R^A_{ai} &:=& [X_i \otimes 1 + 1 \otimes X_i + \delta^k_{ji} \otimes Y^j_k, R^A_a] + \delta^k_{ji} \otimes [Y^j_k,\delta^A_a]~. \label{R} \end{eqnarray} For $n=1$ we get $R^A_i = \delta^k_{ji} \otimes [Y^j_k, \delta^A] \in {\cal H}^0_0 \otimes {\cal H}^0_0$. The Lemma follows from the fact that the commutator with $Y^j_k$ preserves ${\cal H}^0_m$ whereas the commutator with $X_i$ sends elements of ${\cal H}^0_m$ to elements of ${\cal H}^0_{m+1}$. \qed \vskip 2ex \noindent For example, we obtain from (\ref{Ydel}) immediately \begin{equation} \Delta(\delta^k_{ji,l})=\delta^k_{ji,l} \otimes 1 + 1 \otimes \delta^k_{ji,l} + \delta^a_{jl} \otimes \delta^k_{ai} + \delta^a_{il} \otimes \delta^k_{ja} - \delta^k_{al} \otimes \delta^a_{ji}~. \label{d2} \end{equation} The counit $\epsilon$ on ${\cal H}$ is defined by \begin{equation} \varepsilon(1)=1~,\qquad \varepsilon(h)=0\quad \forall h\neq 1 ~. \end{equation} The counit axiom \[ (\varepsilon \otimes {\rm id}) \circ \Delta(h) = ({\rm id} \otimes \varepsilon) \circ \Delta(h) = h\qquad \forall\, h\in {\cal H} \] is clear for $h=X_i,Y^k_j,\delta^A$. For $\delta^A_a$ it follows from Lemma \ref{dex}, using $\varepsilon (h^0)=0$ for $h^0 \in {\cal H}_n^0$. Therefore, ${\cal H}$ is a bialgebra (algebra+coalgebra+compatibility), and our next task is to show the existence of an antipode $S$ on ${\cal H}$, making ${\cal H}$ to a Hopf algebra. The antipode has to satisfy the axioms \begin{eqnarray} S(h_1 h_2) = S(h_2) S(h_1)~, \nonumber \\ m \circ (S \otimes {\rm id}) \circ \Delta(h) = \varepsilon(h)~, \label{Sax} \\ \nonumber m \circ ({\rm id} \otimes S) \circ \Delta(h) = \varepsilon(h)~, \end{eqnarray} for $h,h_1,h_2 \in {\cal H}$, and where $m$ denotes the multiplication. Applying (\ref{Sax}) to $1,Y^j_k,\delta^k_{ji},X_i \in {\cal H}$, in that order, we get \begin{eqnarray} S(1) &=& 1 ~, \nonumber \\ S(Y^j_k) &=& - Y^j_k~, \nonumber \\ S(\delta^k_{ji}) &=& - \delta^k_{ji}~, \\ S(X_i) &=& - X_i + \delta^k_{ji} Y^j_k~, \nonumber \end{eqnarray} The antipode on $\delta^A_a$ is obtained from (\ref{Sax}) by recursion in $|a|$, with the task to prove that the tree possible definitions coincide. First, employing the Sweedler notation $\Delta(R_a)=R_{a(1)} \otimes R_{a(2)}$ (and omitting the summation sign), we have with (\ref{R}) \begin{eqnarray} S(\delta^A_{ai}) &=& -\delta^A_{ai} - m \circ (S \otimes {\rm id}) \circ \Delta(R^A_{ai}) \nonumber \\* &=& -\delta^A_{ai} - S ([X_i,R^A_{a(1)}]) \,R^A_{a(2)} - S (R^A_{a(1)}) \,[X_i, R^A_{a(2)}] - S (\delta^k_{ji} R^A_{a(1)}) \,[Y^j_k, R^A_{a(2)}] \nonumber \\* && {}\qquad - S(\delta^k_{ji}) [Y^j_k,\delta^A_a] \nonumber \\ &=& -\delta^A_{ai} - [X_i-\delta^k_{ji} Y^j_k,S(R^A_{a(1)})] \, R^A_{a(2)} - S (R^A_{a(1)}) \,[X_i, R^A_{a(2)}] \nonumber \\ && {}\qquad + \delta^k_{ji} S(R^A_{a(1)}) \,[Y^j_k, R^A_{a(2)}] + \delta^k_{ji} [Y^j_k,\delta^A_a] \nonumber \\ &=& -\delta^A_{ai} + [-X_i + \delta^k_{ji} Y^j_k, S (R^A_{a(1)}) R^A_{a(2)}] + \delta^k_{ji} [Y^j_k,\delta^A_a] \nonumber \\ &=& [S(\delta^A_a),-X_i+\delta^k_{ji} Y^j_k]=[S(\delta^A_a),S(X_i)]~. \end{eqnarray} In the same way one checks $-\delta^A_{ai} - m \circ ({\rm id} \otimes S) \circ \Delta (R^A_{ai}) = [S(\delta^A_a),S(X_i)]$. For example, one easily obtains \begin{equation} S(\delta^k_{ji,l}) = - \delta^k_{ji,l} + \delta^a_{jl} \delta^k_{ai} + \delta^a_{il} \delta^k_{ja} - \delta^k_{al} \delta^a_{ji}~. \label{Sd2} \end{equation} This finishes our review of the construction of the Connes--Moscovici Hopf algebra \cite{cm2}. In their work, the cyclic cohomology of this Hopf algebra serves as an organizing principle for the computation of the cocycles in the local index formula \cite{cm1}. We hope to be more specific on that point in the future. \section{Explicit solution: rooted trees} Following an idea by Connes and Kreimer \cite{ck} we will now describe the commutative Hopf algebra ${\cal H}_n$ of polynomials in $\delta^A_a$, $|a| \leq n$, by graphical tools, generalized from the one-dimensional case in \cite{ck} to arbitrary dimension of the manifold $M$. In this way we obtain a Hopf algebra of rooted trees, which is intimately related to a Hopf algebra structure in perturbative quantum field theories as discovered by Kreimer \cite{k1}. The antipode of Kreimer's Hopf algebra achieves the renormalization of divergent Feynman graphs, see \cite{k1,k2}. We label the generator $\delta^k_{ji}$ by an indexed dot, \begin{equation} \delta^k_{ji} = ~ \bullet~^k_{ji}~. \end{equation} The goal is to derive the symbol for $\delta^k_{ji,l}$. This goes via the coproduct (\ref{d2}), which tells us after comparison with (\ref{R}) \begin{equation} \bullet~^a_{bl} \otimes [Y^b_a, ~\bullet~^k_{ji}] = {} ~\bullet~^a_{jl} \otimes \bullet~^k_{ai} ~ + ~\bullet~^a_{il} \otimes \bullet~^k_{ja}~ - ~\bullet~^k_{bl} \otimes \bullet~^b_{ji} ~. \label{b2} \end{equation} The commutator with $Y$ picks up one index of $~\bullet~^k_{ji}$ and moves it to the first upper or lower place in $~\bullet~^a_{bl}~$, overwriting the index there. The vacant position in $~\bullet~^k_{ji}$ is filled with the remaining summation index of $~\bullet~^a_{bl}~$. If the index picked up was a lower (upper) one, we count the resulting tensor product positive (negative). This leads us to think of the rhs of (\ref{b2}) as being produced by a cut of a symbol \[ \delta^k_{ji,l} = \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(1,1){$\bullet~_l$} \end{picture}} \quad \longrightarrow \quad \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(0,4){---} \put(1,1){$\bullet~_l$} \end{picture}} = \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ai}$} \put(1,1){$\bullet~^a_{jl}$} \end{picture}} + \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ja}$} \put(1,1){$\bullet~^a_{il}$} \end{picture}} - \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^a_{ij}$} \put(1,1){$\bullet~^k_{al}$} \end{picture}} ~. \] We call the uppermost index which is different from the lower index the root. The graph above the cut connected with the root is called the trunk and goes to the rhs of the tensor product. A graph below the cut is called a cut branch and goes to the lhs of the tensor product. We define the action of a cut as the movement of one index of the vertex above the cut to the first position of the new root of the cut branch. The remaining position to complete the root of the cut branch is filled with a summation index and the same summation index is put into the vacant position of the trunk. In the case of cutting immediately below the root, we have to sum over the three possibilities of picking up indices of the root, adding a minus sign if we pick up the unique upper index. We thus get the following graphical interpretation of (\ref{d2}): \begin{equation} \Delta \left( \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(1,1){$\bullet~_l$} \end{picture}} \right) = \left[ \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(1,1){$\bullet~_l$} \end{picture}} \right]^c + \left[ \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(1,1){$\bullet~_l$} \end{picture}} \right]_c + ~ \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(0,4){---} \put(1,1){$\bullet~_l$} \end{picture}} ~. \label{db2} \end{equation} On the rhs, $[\delta]^c$ stands for $\delta \otimes 1$ (cutting above the entire tree) and $[\delta]_c$ for $1 \otimes \delta$ (cutting below the entire tree). The next step is to compute $\Delta(\delta^k_{ji,lm})$ by commuting $\Delta(X_m)$ with (\ref{db2}). The term $[\delta^k_{ji,l}]^c$ has a non-vanishing commutator only with $X_m \otimes 1$. It yields $\delta^k_{ji,lm} \otimes 1$, and this trivial behavior continues to higher degrees. Next, $X_m \otimes 1$ commutes with $[\delta]_c$, whereas \begin{equation} \left[ X_m \otimes 1 , \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(0,4){---} \put(1,1){$\bullet~_l$} \end{picture}} \right] = \parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(0,11){---} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}} = \delta^a_{jl,m} \otimes \delta^k_{ai} + \delta^a_{il,m} \otimes \delta^k_{ja} - \delta^k_{al,m} \otimes \delta^a_{ji} ~. \label{lm} \end{equation} Our previous definition of a cut extends without modification to that case. The term $1 \otimes X_m$ commuted with $[\delta^k_{ji,l}]_c$ gives $[\delta^k_{ji,lm}]_c$, whereas \begin{equation} \left[ 1 \otimes X_m , \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(0,4){---} \put(1,1){$\bullet~_l$} \end{picture}} \right] = \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1.5,4){---} \put(1,1){$\bullet~_l$} \put(8,1){$\bullet~_m$} \end{picture}} = \delta^a_{jl} \otimes \delta^k_{ai,m} + \delta^a_{il} \otimes \delta^k_{ja,m} - \delta^k_{al} \otimes \delta^a_{ji,m} ~. \end{equation} The cut on the tree in the middle only sees the indices $k,j,i$ -- but not $m$ -- by the definition of a cut as affecting only the indices of the unique vertex above the cut. With this rule we get easily the corresponding expression in terms of $\delta$'s on the rhs. The commutator of $\delta^c_{bm} \otimes Y^b_c$ with $[\delta^k_{ji,l}]_c$ moves the indices $k,j,i,l$ to their correct position in $\delta^c_{bm}$, and this is precisely obtained as the sum of two different cuts: \begin{eqnarray} \left[ \delta^c_{bm} \otimes Y^b_c, \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(1,1){$\bullet~_l$} \end{picture}} \right] &=& \parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(0,4){---} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}} + \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(5.5,4){---} \put(1,1){$\bullet~_l$} \put(8,1){$\bullet~_m$} \end{picture}} ~, \label{lm2} \\[-2ex] \nonumber \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(5.5,4){---} \put(1,1){$\bullet~_l$} \put(8,1){$\bullet~_m$} \end{picture}} &=& \delta^a_{jm} \otimes \delta^k_{ai,l} + \delta^a_{im} \otimes \delta^k_{ja,l} - \delta^k_{am} \otimes \delta^a_{ji,l}~,\quad \parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(0,4){---} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}} = \delta^a_{lm} \otimes \delta^k_{ji,a} ~. \end{eqnarray} There remains one final commutator to compute, that of $\delta^c_{bm} \otimes Y^b_c$ with the graph in (\ref{db2}) already cut. For each of the tree terms corresponding to the previous cut, we have to move each of the tree indices of its root down to $\delta^c_{bm}$. This gives the following symbolic expression of these nine tensor products: \begin{eqnarray} \left[ \delta^c_{bm} \otimes Y^b_c, \parbox{8mm}{\begin{picture}(6,11) \put(1,8){$\bullet~^k_{ji}$} \put(2,9){\line(0,-1){7}} \put(0,4){---} \put(1,1){$\bullet~_l$} \end{picture}} \right] = \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1.5,4){---} \put(6,3){---} \put(1,1){$\bullet~_l$} \put(8,1){$\bullet~_m$} \end{picture}} &=& \delta^a_{jl} \delta^b_{am} \otimes \delta^k_{bi} + \delta^a_{jl} \delta^b_{im} \otimes \delta^k_{ab} - \delta^a_{jl} \delta^k_{bm} \otimes \delta^b_{ai} \nonumber \\[-1ex] &+& \delta^a_{il} \delta^b_{jm} \otimes \delta^k_{ba} + \delta^a_{il} \delta^b_{am} \otimes \delta^k_{jb} - \delta^a_{il} \delta^k_{bm} \otimes \delta^b_{ja} \nonumber \\[0.5ex] &-& \delta^k_{al} \delta^b_{jm} \otimes \delta^a_{bi} - \delta^k_{al} \delta^b_{im} \otimes \delta^a_{jb} + \delta^k_{al} \delta^a_{bm} \otimes \delta^b_{ji} ~.\quad{} \label{cut2} \end{eqnarray} Note that the order of the cuts in this graph is important, we first have to cut the vertex $l$ away and then the vertex $m$. Our construction leads us to define \begin{equation} \delta^k_{ji,lm} = \parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}} + \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1,1){$\bullet~_l$} \put(8,1){$\bullet~_m$} \end{picture}} ~. \end{equation} \begin{Definition} Let $\delta^A_a = \sum_{k=1}^{|a|!} t^{|a|}_k$ be recursively represented by a sum of $|a|!$ connected rooted trees, each of them having $|a|{+}1$ vertices. We define \begin{equation} \delta^A_{ai} \equiv [X_i,\delta^A_a] = \sum_{k=1}^{|a|!} \sum_{j=1}^{|a|+1} t^{|a|}_{k_j} =: \sum_{\ell=1}^{|ai|!} t^{|ai|}_{\ell}~, \label{tkj} \end{equation} where the rooted tree $t^{|a|}_{k_j}$ is obtained by attaching the new vertex $i$ to the right of the $j^{\rm th}$ vertex of $t^{|a|}_k$. \end{Definition} \begin{Proposition} The coproduct of $\delta^A_a = \sum_{k=1}^{|a|!} t^{|a|}_k$ is given by \begin{equation} \Delta (\delta^A_a)= \delta^A_a \otimes 1+1 \otimes \delta^A_a + \sum_{k=1}^{|a|!} \sum_{{\cal C}} P^{\cal C}(t^{|a|}_k) \otimes R^{\cal C}(t^{|a|}_k)~, \label{PR} \end{equation} where for each $t^{|a|}_k$ the sum is over all admissible cuts ${\cal C}$ of $t^{|a|}_k$ (i.e.\ those non-empty multiple cuts for which on each path from the bottom to the root there is at most one individual cut). In eq.\ (\ref{PR}), $R^{\cal C}(t^{|a|}_k)$ is the trunk and $P^{\cal C}(t^{|a|}_k)$ the product of cut branches obtained by cutting $t^{|a|}_k$ via the multiple cut ${\cal C}$. If immediately below a vertex there are several cuts on outgoing edges, the order of the cuts is from left to right. \end{Proposition} {\it Proof.} Commuting $\Delta(X_i)$ with $\Delta(\delta^A_a)$ to get $\Delta(\delta^A_{ai})$, the term $\delta^A_a \otimes 1$ develops into $\delta^A_{ai} \otimes 1$. Next, $X_i \otimes 1$ attaches successively a vertex $i$ to each vertex of the cut branches $P^{\cal C}(t^{|a|}_k)$, and $1 \otimes X_i$ does the same for the trunk $R^{\cal C}(t^{|a|}_k)$ of each tree $t^{|a|}_k$ constituting $\delta^A_a$. Finally $\delta \otimes Y$ attaches a cut-away vertex everywhere on the trunk, not on the cut branch. This excludes multiple cuts on paths from bottom to top. The result clearly reproduces our prescription of the coproduct of $\delta^A_{ai}$, see (\ref{tkj}). \qed \vskip 2ex We make one important observation. Although the operators $\delta$ are invariant under permutation of the indices after the comma, for instance $\delta^k_{ji,lm}=\delta^k_{ji,ml}$, see (\ref{flat}), this symmetry is lost on the level of individual trees, see for instance (\ref{lm}). However, these terms combined with the `diagonal' terms of (\ref{lm2}) are symmetric in $l$ and $m$. We recall that in Kreimer's Hopf algebra of renormalization \cite{k1,k2} a rooted tree represents the divergence structure of a Feynman graph. A divergent sector in such a graph is represented by a vertex. The root represents the overall (superficial) divergence. The construction rule for the tree is -- in absence of overlapping subdivergences -- to put subdivergences $\gamma_i$ of a divergence $\gamma$ into down-going branches of $\gamma$. Disjoint divergences are only indirectly connected via the divergence which contains them as subdivergences. Overlapping divergences have to be resolved in terms of disjoint and nested ones and give a sum of trees, see \cite{kw,k4}. The $n$-dimensional case treated here is closer to quantum field theory than dimension $1$ because we obtain \emph{decorated} trees -- the decoration here being given by spacetime indices (three for the root) whereas in QFT it is a label for divergent Feynman graphs without subdivergences. In this sense, a (not super-) renormalizable QFT has something to do with diffeomorphisms on an infinite dimensional manifold. Our observation leads us to speculate that {\it the sum of Feynman graphs according to the collection of rooted trees to $\delta$'s has more symmetry than the individual Feynman graphs.} This should be checked in QFT calculations. Another interpretation would be the observation \begin{equation} \parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}} + \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1,1){$\bullet~_l$} \put(8,1){$\bullet~_m$} \end{picture}} - \parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(1,8){$\bullet~_m$} \put(1,1){$\bullet~_l$} \end{picture}} - \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1,1){$\bullet~_m$} \put(8,1){$\bullet~_l$} \end{picture}} =0 ~, \label{rel} \end{equation} which could possibly be regarded as a relation between Feynman graphs similar to those derived in \cite{k3}\footnote{Dirk Kreimer confirmed to me that (\ref{rel}) is satisfied in QFT for the leading divergences, as it can be derived from sec.\ V.C in \cite{kd}. For non-leading singularities there will be (probably systematic) modifications.}. \begin{Proposition} The antipode $S$ of $\delta^A_a=\sum_{k=1}^{|a|!} t^{|a|}_k$ is given by \begin{equation} S(\delta^A_a) = - \delta^A_a - \sum_{k=1}^{|a|!} \sum_{{\cal C}_a} (-1)^{|{\cal C}_a|}\; P^{{\cal C}_a}(t^{|a|}_k)\,R^{{\cal C}_a}(t^{|a|}_k)~, \label{Sprp} \end{equation} where the sum is over the set of all non-empty multiple cuts ${\cal C}_a$ of $t^{|a|}_k$ (multiple cuts on paths from bottom to the root are allowed) consisting of $|{\cal C}_a|$ individual cuts. The order of cuts is from top to bottom and from left to right. \end{Proposition} {\it Proof.} We apply the antipode axiom $m \circ(S \otimes {\rm id}) \circ \Delta = 0$, see (\ref{Sax}), to (\ref{PR}), giving with $S(1)=1$ the recursion \[ S(\delta^A_a) = - \delta^A_a - \sum_{k=1}^{|a|!} \sum_{\cal C} \Big( \prod_{j=1}^{|{\cal C}|} S( t^{|a|,{\cal C}}_{k,j}) \Big)\, R^{\cal C}(t^{|a|}_k)~, \qquad P^{\cal C}(t^{|a|}_k)= \prod_{j=1}^{|{\cal C}|} t^{|a|,{\cal C}}_{k,j}~, \] where $|{\cal C}|$ is the number of individual cuts in ${\cal C}$. For each $\{{\cal C},j\}$ we have \begin{equation} S( t^{|a|,{\cal C}}_{k,j})= - t^{|a|,{\cal C}}_{k,j} - \sum_{{\cal C}_j} S(P^{{\cal C}_j}(t^{|a|,{\cal C}}_{k,j})) R^{{\cal C}_j}(t^{|a|,{\cal C}}_{k,j})~, \label{Cj} \end{equation} where the sum is over the set of admissible cuts ${\cal C}_j$ of $t^{|a|,{\cal C}}_{k,j}$. In the first level, the product $\prod_{j=1}^{|{\cal C}|} (- t^{|a|,{\cal C}}_{k,j})$ gives precisely $(-1)^{|{\cal C}|}\; P^{{\cal C}}(t^{|a|}_k)\,R^{{\cal C}}(t^{|a|}_k)$ in (\ref{Sprp}). In the next level, each ${\cal C}_j$ in (\ref{Cj}) leads to a double cut on a path from some bottom vertex in $t^{|a|,{\cal C}}_{k,j}$ to the root of $t^{|a|}_k$, and all double cuts on paths from bottom to root of $t^{|a|}_k$ are obtained (precisely once) in this way. The second cut is below the first one so that the order of cuts is from top to bottom (and from left to right anyway). By recursion one gets all possible cuts ${\cal C}_a$ of $t^{|a|}_k$ contributing with the sign $(-1)^{|{\cal C}_a|}$ to the antipode. \qed \vskip 1ex For $\delta^k_{ji,lm}$, the prescription (\ref{Sprp}) leads to the following antipode: \begin{eqnarray*} S(\delta^k_{ji,lm}) &=& - \parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}} - \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1,1){$\bullet~_l$} \put(8,1){$\bullet~_m$} \end{picture}} \quad \longrightarrow \quad - \delta^k_{ji,lm} \\ &+& \parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(0,4){---} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}} \quad \longrightarrow \quad + \delta^a_{lm} \delta^k_{ji,a} \\ &+&\parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(0,11){---} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}}\quad \longrightarrow \quad + \delta^a_{jl,m} \delta^k_{ai} + \delta^a_{il,m} \delta^k_{ja} - \delta^k_{al,m} \delta^a_{ji} \\ &-&\parbox{8mm}{\begin{picture}(6,18) \put(1,15){$\bullet~^k_{ji}$} \put(2,16){\line(0,-1){14}} \put(0,11){---} \put(0,4){---} \put(1,8){$\bullet~_l$} \put(1,1){$\bullet~_m$} \end{picture}}\quad \longrightarrow \quad \begin{array}[t]{l} - \delta^b_{jm} \delta^a_{bl} \delta^k_{ai} - \delta^b_{lm} \delta^a_{jb} \delta^k_{ai} + \delta^a_{bm} \delta^b_{jl} \delta^k_{ai} \\ - \delta^b_{im} \delta^a_{bl} \delta^k_{ja} - \delta^b_{lm} \delta^a_{ib} \delta^k_{ja} + \delta^a_{bm} \delta^b_{il} \delta^k_{ja} \\ + \delta^b_{am} \delta^k_{bl} \delta^a_{ji} + \delta^b_{lm} \delta^k_{ab} \delta^a_{ji} - \delta^k_{bm} \delta^b_{al} \delta^a_{ji}\end{array} \\ &+& \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1,1){$\bullet~_l$} \put(1.5,4){---} \put(8,1){$\bullet~_m$} \end{picture}} \quad \longrightarrow \quad + \delta^a_{jl} \delta^k_{ai,m} + \delta^a_{il} \delta^k_{ja,m} - \delta^k_{al} \delta^a_{ji,m} \\ &+& \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1,1){$\bullet~_l$} \put(5.5,4){---} \put(8,1){$\bullet~_m$} \end{picture}} \quad \longrightarrow \quad + \delta^a_{jm} \delta^k_{ai,l} + \delta^a_{im} \delta^k_{ja,l} - \delta^k_{am} \delta^a_{ji,l} \\ &-& \parbox{14mm}{\begin{picture}(14,11) \put(4.5,8){$\bullet~^k_{ji}$} \put(5.5,9){\line(-1,-2){3.5}} \put(5.5,9){\line(1,-2){3.5}} \put(1,1){$\bullet~_l$} \put(1.5,4){---} \put(6,3){---} \put(8,1){$\bullet~_m$} \end{picture}} \quad \longrightarrow \quad \begin{array}[t]{l} - \delta^a_{jl} \delta^b_{am} \delta^k_{bi} - \delta^a_{jl} \delta^b_{im} \delta^k_{ab} + \delta^a_{jl} \delta^k_{bm} \delta^b_{ai} \\ - \delta^a_{il} \delta^b_{jm} \delta^k_{ba} - \delta^a_{il} \delta^b_{am} \delta^k_{jb} + \delta^a_{il} \delta^k_{bm} \delta^b_{ja} \\[0.5ex] + \delta^k_{al} \delta^b_{jm} \delta^a_{bi} + \delta^k_{al} \delta^b_{im} \delta^a_{jb} - \delta^k_{al} \delta^a_{bm} \delta^b_{ji} \end{array} \end{eqnarray*} One checks, using (\ref{Sd2}) and (\ref{lm})--(\ref{cut2}), the antipode axioms (\ref{Sax}). \section*{Acknowledgements} First of all I would like to thank Alain Connes for his help on a point were I was stuck. I profited very much from seminar notes by Daniel Testard especially on the passage to the flat case. I hope to include this in a future version. It is a pleasure to thank Bruno Iochum, Thomas Krajewski, Serge Lazzarini, Thomas Sch\"ucker and Daniel Testard for encouragement and numerous discussions. I am grateful to Dirk Kreimer for a comment on eq.\ (\ref{rel}).